You are on page 1of 681

Advances in Rockfill Structures

NATO ASI Series


Advanced Science Institutes Series

A Series presenting the results of activities sponsored by the NA TO Science Committee,


which aims at the dissemination of advanced scientific and technological knowledge,
with a view to strengthening links between scientific communities.

The Series is published by an international board of publishers in conjunction with the


NATO Scientific Affairs Division

A Life Sciences Plenum Publishing Corporation


B Physics London and New York

C Mathematical Kluwer Academic Publishers


and Physical Sciences Dordrecht, Boston and London
D Behavioural and Social Sciences
E Applied Sciences

F Computer and Systems Sciences Springer-Verlag


G Ecological Sciences Berlin, Heidelberg, New York, London,
H Cell Biology Paris and Tokyo
I Global Environmental Change

NATo-PCO-DATA BASE

The electronic index to the NATO ASI Series provides full bibliographical references
(with keywords and/or abstracts) to more than 30000 contributions from international
scientists published in all sections of the NATO ASI Series.
Access to the NATO-peO-DATA BASE is possible in two ways:
- via online FILE 128 (NATO-PCO-DATA BASE) hosted by ESRIN,
Via Galileo Galilei, 1-00044 Frascati, Italy.
- via CD-ROM "NATO-PCO-DATA BASE" with user-friendly retrieval software in
English, French and German ( WTV GmbH and DATAWARE Technologies Inc.
1989).

The CD-ROM can be ordered through any member of the Board of Publishers or
through NATO-PCO, Overijse, Belgium.

Series E: Applied Sciences - Vol. 200


Advances in
Rockfill Structures
edited by

E. Maranha das Neves


Geotechnical Department,
National Laboratory of Civil Engineering,
Lisbon, Portugal

"
Iit.AI

SPRINGER-SCIENCE+BUSINESS MEDIA, B.V.


Based on the NATO Advanced Study Institute on
Advances in Rockfill Structures
Lisbon, Portugal
18-29 June 1990

Library of Congress Cataloging-in-Publication Data

NATCl Advanced Study Inst i tute on Advances in Rockf i II Structures (1990


NAT[)
l1~t'on, Portugal)
Aavances in rocktl
rockt' II structures! eOlted by E, Maranha das Neves,
structures ! eo'ted
p, cm, -- (NATO ASI ser,es, serles, Series E, Applied sciences;
sciences ; v,
:"00)
"Proceedings of the NATO Advanced Study Inst'tute Instltute on Advances ,n
ln
Rockfi II Structures, held ,n ln L'sbon, Portugal, 18-29 June, 1990,"
Inc II udes b,
b 1 b li
Ii ograph i ca II references,
ISBN 978-94-010-5414-0 ISBN 978-94-011-3206-0 (eBook)
DOI 10.1007/978-94-011-3206-0
1, Rockfills--Congresses, 2, Earth dams--Congresses, I, Neves,
E, Maranha das (Emanuel Maranha), 1938- II. Tltle,
Ir. THle,
III, Ser,es:
111, Serles: NATO ASI series. Series E, Applied sciences no, 200.
7A709."37 1990
624.1'832--dc20 91-14610

ISBN 978-94-010-5414-0

Printed on acid-free paper

All Rights Reserved


1991 Springer Science+Business Media Dordrecht
Originally published by Kluwer Academic Publishers in 1991
Softcover reprint ofthe
of the hardcover 1st edition 1991
No part of the material protected by this copyright notice may be reproduced or
utilized in any form or by any means, electronic er or mechanical, including photo-
copying, recording or by any information storage and retrieval system, without written
ccpying,
permission from the copyright owner.
CONTENTS

LIST OF CONTRIBUTORS XXI


FOREWORD XXIII
CHAPTER 1
ROCKFILL STRUCTURES: THE PRESENT AND THE FUTURE
E. MARANHA das NEVES 1

CHAPTER 2
PHYSICAL CHARACTERIZATION AND ASSESSMENT OF ROCK
DURABILITY THROUGH INDEX PROPERTIES
J. DELGADO RODRIGUES

1. INTRODUCTION 7
2. BRIEF PRESENTATION OF MOST COMMON ROCK TYPES 8
2.1. significance of geological classifications 8
2.2. Rock materials in general classifications 9

3. PROPERTIES OF ROCK MATERIALS 10


3.1. General 10
3.2. Rock masses and rock materials 11
3.3. Brief considerations about sampling 11

4. LABORATORY CHARACTERIZATION OF ROCK MATERIALS 13


4.1. General 13
4.2. Intrinsic properties 14
4.3. Index properties 15
4.3.1. General 15
4.3.2. Some common index properties 15
4.3.3. Estimation of rockfill character-
istics through index properties 20
4.4. Shape and size properties 22
4.5. Durability 23
4.5.1. General remarks 23
4.5.2. Some methods of rock durability
assessment 24
4.5.3. Assessment of rock durability
through index properties 25
VI

CHAPTER 3
ROCKFILL MODELLING
A. K. PARKIN

1. INTRODUCTION 35

2. ROCKFILL IN PLACE 36
3. TRIAXIAL TESTING EQUIPMENT 36
4. MAXIMUM PARTICLE SIZE, TEST SAMPLES 36

5. MODEL GRADINGS 39

6. SOME PARTICULAR ISSUES RELATING TO THE


OEDOMETER TEST 40

6.1. Oedometer dimensions 40


6.2. Side friction models 42
6.3. Effects of initial stress 45

7. A CASE HISTORY 46

8. CONCLUSIONS 48

ACKNOWLEDGEMENT 49

REFERENCES 49

CHAPTER 4
LABORATORY SHEAR STRENGTH TESTS AND THE STABILITY OF
ROCKFILL SLOPES
J. A. CHARLES

1. INTRODUCTION 53

2. LABORATORY TESTS 54

3. STRAIN CONDITIONS 54

4. INITIAL POROSITY 55

5. CONFINING PRESSURE 55
vii

5.1. Curved failure envelope 56


5.2. Relationship between ~' and cr' 57

6. DILATANCY 60

6.1. A basic angle of shearing resistance 60


6.2. A component due to dilatancy 60

7. PARTICLE SIZE 60

8. SLOPE STABILITY 62
9. STABILITY CHARTS FOR ROCKFILL SLOPES 63
10. STABILITY OF SUBMERGED ROCKFILL SLOPES 64

11. DESIGN OF ROCKFILL SLOPES 67

11.1. Determination of the rockfill


shear strength parameters 67
11.2. Selection of an appropriate factor
of safety 67
11.3. Calculation of the magnitude of the
stability number 69
11.4. Determination of the slope angle B 69

12. EXAMPLES OF USE OF STABILITY CHARTS 69

12.1. Example; 100 m high embankment 69


12.2. Example; 10 m high embankment 69
12.3. Discussion 69

13. CONCLUDING REMARKS 70

ACKNOWLEDGEMENT 70

REFERENCES 70

CHAPTER 5
LABORATORY COMPRESSION TESTS AND THE DEFORMATION OF
ROCKFILL STRUCTURES
J. A. CHARLES

1. INTRODUCTION 73
1.1. Changes in applied stress 73
1.2. Increase in moisture content 74
1.3. Vibrations associated with dynamic
loading 74
viii

2. LABORATORY ONE DIMENSIONAL COMPRESSION TESTING 74

3. FIELD PROPERTIES 80

4. FIELD MONITORING OF DEFORMATIONS 81

5. CONSTRUCTION DEFORMATIONS OF EMBANKMENTS 81

6. MOVEMENT OF UPSTREAM MEMBRANE EMBANKMENT


DAMS DUE TO RESERVOIR IMPOUNDING 83

7. MOVEMENT OF CENTRAL CORE EMBANKMENT DAMS


DUE TO RESERVOIR IMPOUNDING 87

8. MOVEMENT OF ROCKFILL STRUCTURES DUE TO


COLLAPSE COMPRESSION 87

9. CREEP SETTLEMENT OF ROCKFILL STRUCTURES 89

10. CONCLUDING REMARKS 92

ACKNOWLEDGEMENTS 92

APPENDIX A
CONSTANT EQUIVALENT CONSTRAINED MODULUS 93

APPENDIX B
STRESS PATHS DURING CONSTRUCTION AND RESER-
VOIR IMPOUNDING FOR UPSTREAM MEMBRANE DAMS 93

REFERENCES 94

CHAPTER 6
COLLAPSE: ITS IMPORTANCE, FUNDAMENTALS AND
MODELLING
J. L. JUSTO

1. INTRODUCTION 97

2. THE FUNDAMENTALS OF COLLAPSE IN ROCKFILL 99

3. COLLAPSE MODELLING 100


4. ONE-DIMENSIONAL COLLAPSE DURING WATER RISE
IN A GRANULAR MATERIAL. BUOYANCY AND CREEP 121

5. COLLAPSE PRODUCED CRACKS 127

6. COLLAPSE AND POST-CONSTRUCTIVE SETTLEMENTS


OF ROCKFILL DAMS 132
ix

6.1. Central core dams 132


6.2. Martin Gonzalo Rockfill dam 135
6.3. Post-Constructive settlements of
rockfill dams 137
7. CONCLUSIONS 140

REFERENCES 141
APPENDIX 1: PARAMETERS OF CONGLOMERATE IN
YEGUAS DAM, ASSUMING OEDOMETRIC
CONDITIONS (FIG.25 AND 26) 143
APPENDIX 2: PARAMETERS OF ROCKFILL IN MARTIN
GONZALO DAM, ASSUMING OEDOMETRIC
CONDITIONS 146

CHAPTER 7
TEST FILLS AND IN SITU TESTS
J. L. JUSTO

1. WHY IN SITU TESTS? 153


2. TEST FILLS 154
3. PLATE LOADING TESTS 158
4. IN SITU DENSITY 167
5. PERMEABILITY TEST 170

6. TENSION TESTS (URIEL AND PEREZ, 1981) 176


7. SHEAR STRENGTH TESTS 178
8. CONCLUSIONS 182
REFERENCES 188
APPENDIX 1: INFILTRATION FROM A SHALLOW
EXCAVATION 190

APPENDIX 2: INFILTRATION FROM CASED HOLES 193


x

CHAPTER 8
LABORATORY TESTING AND QUALITY CONTROL OF ROCKFILL
- GERMAN PRACTICE
J. BRAUNS AND K. KAST

1. INTRODUCTION 195
2. GENERAL ASPECTS 195
3. ASPECTS OF LABORATORY TESTING 204
3.1. Rock quality and gradation 204
3.2. "True" rockfill samples 204
3.3. Sample dimensions 207
3.4. Layout of devices for oedometer tests 207
3.5. Layout of devices for triaxial tests 209
3.6. Direct shear tests 212
4. ASPECTS OF QUALITY CONTROL 213
5. CONCLUDING REMARKS 218
REFERENCES 219

CHAPTER 9
CREEP OF ROCKFILL
A. K. PARKIN

1. INTRODUCTION 221

2. RATE METHODS APPLIED TO SETTLEMENT ANALYSIS 222

3. ROCKFILL CREEP IN OEDOMETER COMPRESSION 224

4. APPLICATION TO FIELD SETTLEMENT RECORDS 225

5. CREST SETTLEMENT OF DAMS 228

6. LOAD TESTS ON LARGE BORED PILES 232

7. CONCLUSIONS 234
ACKNOWLEDGEMENT 236
REFERENCES 236
xi

CHAPTER 10
FILTERS AND DRAINS
J. BRAUNS

1. INTRODUCTION 239
2. PRESENT PRACTICE OF FILTER DESIGN 241
2.1. Geometrical criteria 241
2.2. Scattering of gradations 242
2.3. Finest fraction in filter materials 246
2.4. Filters for cohesive soils 247
2.5. Hydraulic criteria 248
3. RECENT INVESTIGATIONS ON THE PROBLEM OF
FILTER STABILITY 252
4. DRAINS 261
5. CONCLUDING REMARKS 265
REFERENCES 266

CHAPTER 11
STRESS - STRAIN LAWS AND PARAMETER VALUES
D. J. NAYLOR

1. INTRODUCTION 269

2. HYPERBOLIC AND Ec-Ko MODELS 271

2.1. Background 271


2.2. Hyperbolic model: formulation 271
2.3. Ec
-K omodel:
. formulation 272
2.4. Hyperbollc and Ec-Ko model: parameters 273
3. K-G MODEL 274

3.1. Background 274


3.2. Formulation 274
3.3. K-G Parameters 275
3.4. K-G Model - An alternative 275
4. CRITICAL STATE MODEL 279
xii

4.1. Background 279


4.2. Basic formulation 280
4.3. Variations on the theme 285
4.4. C.S. Parameters 286
5. CONCLUSIONS 289
REFERENCES 290

CHAPTER 12
FINITE ELEMENT METHODS FOR FILLS AND EMBANKMENT
DAMS
D. J. NAYLOR

1. INTRODUCTION 291
2. NUMBER OF LAYERS - ACTUAL AND ANALYTICAL 292
3. DEFORMATION IN A RISING FILL 292
4. BASIC FINITE ELEMENT PROCEDURE 292
5. INTERPRETATION OF FINITE ELEMENT DIS-
PLACEMENTS - 1D CASE 294
6. NEW LAYER STIFFNESS REDUCTION 296
7. MODELLING COMPACTION 300
8. FINITE ELEMENT EFFECTIVE STRESS TECHNIQUES 302
8.1. Undrained effective stress analysis 302
8.2. Known pore pressure change analysis 305
9. FIRST FILLING AND OPERATION - GENERAL 306

10. LOADING DUE TO IMPOUNDING 308


10.1. upstream membrane dam 308
10.2. Internal membrane dam 308
10.3. Zoned embankment dams 312
11. ANALYSIS OF FIRST FILLING AND OPERATION 312
11.1. First filling 312
11.2. Steady seepage condition 314
11.3. Finite element considerations 314
12. COLLAPSE SETTLEMENT 314
xili

12.1. Nobari and Duncan's method 317


12.2. Generalisation of Nobari and
Duncan's method 319
12.3. One-dimensional example 320
13. APPLICATIONS 323
13.1. carsington dam 323
13.2. Beliche dam 325
13.3. Monasavu dam 330
REFERENCES 335
APPENDIX: DERIVATION OF EQUIVALENT LAYER
STIFFNESS 332

CHAPTER 13
CONCRETE FACE ROCKFILL DAMS
NELSON L. DE S. PINTO
1. INTRODUCTION 341
2. CURRENT DESIGN PRACTICE 343
2.1. Evolution 343
2.2. Embankment 344
2.2.1. General comments 344
2.2.2. Zone 1. Impervious blanket 345
2.2.3. Zone 2. Processed small rock
transition 345
2.2.4. Zone 3. Main rockfill embankment 347
2.2.5. Fill cross section 348
2.3. Plinth 349
2.4. Concrete face 351
2.4.1. Slab thickness 351
2.4.2. Concrete 351
2.4.3. Reinforcing 351
2.4.4. Joints 351
2.4.5. Joint details 351

3. CONSTRUCTION FEATURES 356

3.1. Embankment 356


3.2. Concrete works 359
3.3. River handling aspects 363
4. MONITORING AND BEHAVIOUR 366
4.1. Dam movements 366
dv

4.2. Performance under seismic load 368


REFERENCES 371

CHAPTER 14
STATIC BEHAVIOUR OF EARTH-ROCKFILL DAMS
E. MARANHA das NEVES
1. INTRODUCTION 375
2. GEOMETRICAL PHYSICAL AND MECHANICAL DATA
OF A SERIES OF EARTH-ROCKFILL DAMS 377
3. STRUCTURAL BEHAVIOUR AND EXPERIENCE 386
3.1. Construction materials 388
3.2. Placement techniques 389
3.3. Structural conception 389
3.3.1. Core slenderness 389
3.3.2. Inclination of the dam slopes 390
3.3.3. Ponti on of the core 391
3.3.4. Deformability of the different
zones of the dam 394
3.3.5. Filters 395
3.3.6. Shape of the valley 409
3.3.7. Configuration in plan 414
3.4. Final remarks 415
4. MODELLING THE STRUCTURAL BEHAVIOUR 416
4.1. Dam with a vertical core 419
4.2. Dam with a sloping core 423
4.3. Influence of the dam height 424
4.4. Deformations 426
4.5. Final remarks 428

5. SAFETY EVALUATION AND THE LIMIT STATES


CONCEPT 428
5.1. Methods for evaluating dam safety 429
5.2. Brief notes about limit states design 429
5.3. Overall safety factor and probability
of failure 431
5.4. Partial safety factors and limit states 432
5.5. Use of partial safety factors in safety
analysis of an earth-rockfill dam 435
5.5.1. Linear elastic model 435
5.5.2. Non-linear elastic model 437
5.6. Final remarks 439
xv

6. CONCLUSIONS 439

ACKNOWLEDGEMENTS 440

REFERENCES 441

CHAPTER 15
DYNAMIC BEHAVIOUR OF ROCKFILL DAM
E. YANAGISAWA

1. INTRODUCTION 449

2. DAMAGE TO FILL DAMS DUE TO STRONG


EARTHQUAKES 450

2.1. Damage to fill dams by strong shaking 450


2.2 Soil liquefaction 451

3. DYNAMIC PROPERTIES OF FILL DAM MATERIALS 452

3.1. Dynamic deformation characteristics


of fill dam materials 452
3.2. Dynamic strength of soils 455

4. RESPONSE ANALYSIS OF FILL DAMS 456

4.1. Shear beam theory 456


4.2. Response analyses of rockfill dam 458

5. LIQUEFACTION ANALYSIS 461

5.1. The constitutive equation for un-


drained shear behavior of sands 461
5.2. Pore pressure generated during earth-
quake 463

6. EARTHQUAKE RESISTANT DESIGN OF FILL DAMS


IN JAPAN 466

6.1. Factor of safety 466


6.2. Dynamic analyses 466

7. CONCLUSIONS 467

ACKNOWLEDGEMENT 468

REFERENCES 468
XVI

CHAPTER 16
MONITORING AND SAFETY EVALUATION OF ROCKFILL DAMS
A. VEIGA PINTO

1. INTRODUCTION 471
2. TYPE OF MEASUREMENTS 473
3. MONITORING SCHEME DESIGN 475
3.1. Selection of monitoring equipment 475
3.2. Selection of instruments locations 477
3.3. Installation plans and procedures
subsequent to construction phase 479
3.4. Monitoring frequencies 479
3.5. Plan of first filling 479
4. MONITORING EQUIPMENT 481
4.1. Triangulation and trilateration
networks 482
4.2. Precision levelling 483
4.3. Inclinometer 483
4.4. Fluid level settlement gauge 487
4.5. Horizontal displacements device 488
4.6. Total pressure cells 490
4.7. Piezometer 492
4.8. Seepage monitoring 495
4.9. Earthquake effect monitoring 497
5. READINGS, PROCESSING AND ANALYSIS OF RESULTS 497
5.1. Data collection 497
5.2. Data transmission 499
5.3. Data processing and information storage 499
5.4. Data presentation 500
5.5. Performance evaluation 500

6. VISUAL INSPECTION 501

7. SAFETY EVALUATION BASED ON DETERIORATION 502

7.1. Introduction 502


7.2. statistical analysis 503
7.3. Remedial measures 503
xvii

8. DAM SAFETY REGULATIONS 508


9. STRAINS OBSERVED IN ROCKFILL DAMS 510
9.1. Introduction 510
9.2. Construction phase 511
9.3. After construction 512
9.4. After earthquakes 515
10. CONCLUSIONS 518
REFERENCES 520

CHAPTER 17
PRINCIPLES OF ROCKFILL HYDRAULICS
R. MARTINS
1. INTRODUCTION 523
1.1. Definition of rockfill hydraulics 523
1.2. Complements to the former definition 524
1.3. Scope of rockfill hydraulics 525
1.4. Subjects dealt with in this chapter 525
2. CHARACTERIZATION OF ROCKFILL 526

2.1. Preliminary hypotheses 526


2.2. Size 526
2.3. Shape 527
2.4. Disposition 529
2.5. Specific gravity 530
2.6. Friction angle 531
2.7. Final comments on sources of uncer-
tainty in rockfill hydraulics 532
3. FRICTION HEAD LOSSES IN OPEN CHANNELS 533
3.1. Preliminary remarks 533
3.2. Resistance laws 535
3.3. Function f () in case of high
relative roughness 536
3.4. Data for the use of the Gauckler-
-Manning expression 537
3.5. Conclusions 538
XVllI

4. STABILITY OF ROCKFILL SUBJECT TO FLOW 539


4.1. Preliminary remarks 539
4.2. Stability in bidimensional channels
with horizontal or quasi-horizontal
bed and non-high relative roughness 540
4.3. Case of high relative roughness 542
4.4. Channels with non-horizontal bed 543
4.5. Stability in trapezoidal channels 544
4.6. Stability in bends 545
4.7. Effects of lining thickness, grada-
tion, shape and specific gravity 545
4.8. Conclusions 547
5. SEEPAGE FLOW 549
5.1. Preliminary remarks 549
5.2. Mean hydraulic radius of the voids
and mean velocity in the voids 552
5.3. Turbulent seepage flow 553
5.4. Transition zone 555
5.5. Conclusions 557
ACKNOWLEDGEMENTS 558

REFERENCES 558
ANNEX 1: EXAMPLES OF CALCULATING THE
CARACTERISTIC. DIMENSIONS OF
BLOCK SETS. 564

ANNEX 2: HIDRAULIC GRADIENT 565

ANNEX 3: A REASON FOR APPARENT NON-


-LINEARITY IN LAMINAR SEEPAGE
FLOW (CONCEPTUAL EXAMPLE) 567
ANNEX 4: TESTS FOR THE QUADRATIC ZONE 568
ANNEX 5: COMPARISON OF RESULTS FROM EQ.
35 AND FROM THE EXPRESSIONS OF
WILKINS AND JAIN ET AL. 570

CHAPTER 18
THROUGH AND OVERFLOW ROCKFILL DAMS
A. K. PARKIN
1. INTRODUCTION 571
XIX

2. EARLY DEVELOPMENTS IN FLOOD-PROTECTED


ROCKFILLS 572

3. THE SELF-SPILLWAY (THROUGHFLOW) DAM 573


4. LABORATORY STUDIES 574
4.1. Equations of Flow 575
4.2. Hydraulic control points 576
4.3. Analysis of flow fields 577
4.4. Stability 579
5. OVERFLOW ROCKFILLS 580
6. MESH-PROTECTED ROCKFILLS 583
6.1. Bar spacing and configuration 587
6.2. Performance 587
6.3. Protection of cohesive or impervious
fills 587
6.4. Permanent flood protection 588
7. CONCLUSION 589
ACKNOWLEDGEMENT 590
REFERENCES 590

CHAPTER 19
SPECIFICATIONS AND CONTROL OF NATURAL ROCKFILLS
H. EVRARD

1. INTRODUCTION 593

1.1. The technical context 593


1.2. The economic context 593
2. SCALING ROCKFILL REQUIREMENTS 594

3. RECOMMENDATIONS FOR SPECIFICATIONS 598

3.1. Rockfill density 599


3.2. Rockfill properties 600
4. CONTROL OF THE INTRINSIC PROPERTIES
OF THE ROCK 603
5. INSPECTION AND CONTROL OF SUPPLIES 607
6. PREPARATION CONTROL 607
xx

7. CONCLUSION 608

BIBLIOGRAPHY 609

CHAPl'ER 20
ASPHALTIC CONCRETE FACE DAMS
J. L. JUSTO
l. INTRODUCTION 6ll
2. REVETMENT STRUCTURE 623

3. CONSTRUCTION 625

4. THE DEFORMABILITY OF ASPHALTIC CONCRETE


RELATED TO THE STRAINS SUFFERED BY THE
FACING 629

5. PLINTHS 633

6. FINITE ELEMENT COMPUTATIONS 638

7. UPSTREAM SLOPE 642

8. PERFORMANCE OF ASPHALTiC CONCRETE FACING


ROCKFILL DAMS 642

9. THE FUTURE OF ASPHALTIC CONCRETE FACINGS 646

REFERENCES 648

CWSING SESSICN 65l

LIST OF PARTICIPANTS 657


LIST OF CONTRIBUTORS

A. PARKIN
Senior LeclIIrer, Monash Universiry Clayton, Melbourne, \lic((Jria 3168, Australia.

A. VEIGA PINTO
Senior Research Officer, Lab. Nac. Eng. Civil, Av. do Brasil 101, 1799 Lisboa Codex,
Portl.g,i/.

D. NAYLOR
Senior Leclllrer, University College of Swansea, Depart. of Civil Engineering, Single-
ton Park, Swansea, SA2 8PP, U. K.

E. ivlARANHA das NEVES


Head Geotechnical Department, Lah. Nac. Eng. Civil, Av. do Brasil 101, 1799 Lisboa
Codex, PorIl/gal.

E. YANAGISAWA
Professor, To/wku University, Depart. of Civil Engilleering, FaCility of EnRineerillg,
Aoha, Sendai 980, Japan.

If. EVRARD
Head of the Rock Mechanics Group, Laboratoire Regional des Ponts et Chaussee,
CETE de Lyon ]()9, Avenue Salvador-Allende CSE No.1 - 69674 Broil Cedex, France.

J. BRAUNS
Head of Section of Soil and Rock Mechanics, Karlsruhe University, Post/llcll 6980
D - 7500 Karlsruhe, Germany.

xxi
xxii

J. CHARLES
Geotechnics Division, Building Research Establishment Garston Watford WD2 7JR,
U.K.

1. DELGADO RODRIGUES
Principal Research Officer, Lab. Nac. Eng. Civil, Av. do Brasil 101, 1799 Lisboa
Codex, Portugal.

1. JuSTO ALPANES
Senior Lecturer, E. T. S. Arquitecture, Av. Reina Mercedes sin, 41012 Seville, Spain.

N. SOUSA PINTO
Consulting Engineering, Av. Vicente Machado, 2340 , 80430 Curitiba - PR, Brasil.

R. MARTINS
Principal Research, Lab. Nac. Eng. Civil, Av. do Brasil 101, 1799 Lisboa Codex,
Portugal.
FOREWORD

On 1990 June 18-25, an Advanced study Institute (ASI) on


Rockfill structures was held in Lisbon PORTUGAL, at the
Laboratorio Nacional de Engenharia civil (LNEC), having the
NATO Scientific Affairs Division as main sponsor, and the
LNEC, the Junta de Investiga~ao Cientifica e Tecnologica, and
the Funda~ao do Oriente as co-sponsors.
The objective of this ASI was the discussion and updating of
concepts related to the design, construction, operation, and
monitoring of rockfill structures.
In recent years, an increasing use has been made of
rockfills in the construction industry. This trend results
from the great progress made in all technologies related to
the quarrying, transportation, and placement of rock
materials, from the significant advances in the performance
shown by rockfill structures, and, last but not least, from
the abundance and low cost of the rock materials.
The characteristic problems of rockfill constructions have
been occasionally dealt with at some meetings (conferences,
symposia, workshops, etc.) and in odd chapters of books
devoted to several types of works. It was therefore felt that
the matter should be tackled on an overall basis, covering the
various points of view from which rockfills may be regarded.
The ASI was attended by 57 participants, from 18 different
countries, and the lessons given are the basis of the 20
chapters of this book. A state-of-art of the concerned
subjects has thus been obtained. All these results were only
possible due to the highly esteemed support of the NATO
Scientific Affairs Division which is strongly acknowledged and
thanked.

The organizing Committee


E. Maranha das Neves
(Director)
J. Andrew Charles
J. L. Justo Alpanes
A. Veiga Pinto
xxiii
1 - A. CHARLES
2 - D. NAYLOR
3 - J. ALPANES
4 - E. MARANHA DAS NEVES
5 - R. OLIVEIRA
6 - A. VEIGA PINTO
7 - J. MATEUS DA SILVA
8 - C. QUADROS
9 - B. SIYAHI
10 - A. 9ELEBI
11 - P. SECO E PINTO
12 - M. EMILIA BORRALHO
13 - A. PARKIN
14 - J. BARROS GOMES
15 - J. DELGADO RODRIGUES
16 - J. LOUREIRO
17 - F. FEDERICO
18 - A. CORREIA
19 - A. TAN
20 - F. ALMEIDA
21 - H. YILDIRIM
22 - M. SIYAHI
23 - o. FILHO
24 - L. ALMEIDA
25 - F. LUCAS
26 - RUI MARTINS
27 - M. CEDERSTROM
28 - N. JOHANSSON
29 - DA-MANG LEE
30 - C. SANTOS PEREIRA
31 - D. GUIMARAES
32 - A. SILVA
33 - V. JESUS
34 - L. CARTAXO
35 - L. VIRGEN
36 - J. COUTO MARQUES
37 - J. CAVILHAS
38 - ERNESTO DOMINGUES
39 - N. KOLFF
40 - J. AZANEDO
41 - M. PACHAKIS
42 - J. CORDOVA
43 - M. ZACAS
44 - D. MATTAR JUNIOR
45 - J. MATEUS DE BRITO
46 - A. MOFFAT
47 - E. YANAGISAWA
48 - 1. PYRAH
49 - ARMINDO FERREIRA
50 - Mrs. B. FILHO
51 - MOZART B. FILHO
OPENING SESSION
E. MARANHA das NEVES

As was announced in the first bulletin of this NATO Advanced


study Institute, its aim is the dissemination of advanced
scientific knowledge concerning rockfill structures which has
not found its way into university curricula, and to foster
international contacts among scientists.
This ini tiati ve is the consequence of a great effort as
regards research into rockfills made by LNEC, the results of
which became evident, when practical applications were
envisaged in Portugal. Millions of cubic meters of rockfill
have been used in dams, motorways, airports, harbours and
embankments in general, and certainly in the future we will
see an increasing use of this material, with which Portugal
has been so well provided by Nature.
Rockfill structures are used allover the world, but their
role and importance are not generally recognised. For
instance, when reference is made to dams, the public in
general, immediately thinks in terms of concrete dams,
ignorant of the fact that embankment dams are by far the most
numerous; and the surprise is total when it is said that the
highest dams in the world (more than three hundred meters) are
earth-rockfill dams.
Nevertheless, progress in this area is relatively recent and
is centered in three domains:
- construction technology, where compaction by vibration
has a leading role;

- constitutive laws, subject to which fundamental


research on particulate media, laboratory and field
tests, as well as monitoring, have made important
contributions;
numerical methods, because they allow the application
of constitutive laws, and thus, the forecasting of
rockfill structure behaviour.
xxvii
xxviii

It is important to stress that rockfills, when compared with


soils, present additional diff icul ties, when a theoretical
approach to their behaviour is attempted.
One of the most significant, is that for the range of
stresses found in civil engineering problems, particles of
granular materials undergo important breakage, even for very
low stress levels. This means that when travelling along their
stress paths, rockfills are continuously changing not only the
void ratio - as is the case with soils - but also the grain
size. For each new step in this path, a new material is
obtained. It therefore becomes clear how hard it is to design
a rockfill structure when we aim to tackle all the safety
problems involved. Nevertheless, it is our hope that at the
end of this Course a contribution will have been made towards
a clarification of these problems.
Though this Course is intended primarily for NATO countries,
technicians non-NATO countries may also attend it.
We therefore have lecturers and participants from such
different countries as Angola, Australia, Brazil, Cape Verde,
France, Germany, Greece, Italy, Japan, Mexico, Mozambique,
Spain, Sweden, Turkey, the united Kingdom, the united States
and, of course, Portugal.
The total number of lecturers is twelve, and the number of
participants is about seventy.
The course lasts for two weeks, with forty-two hours of
lectures, and discussions periods devoted to the contributions
of participants; a panel on discussion of the future of
rockfills an editorial meeting, and two technical visits: one
to the LNEC (next Wednesday) and the other, to a section of
the Lisbon-oporto motorway, where rockfills are being used in
road embankments.
The objective of this course is not only to contribute to
the advancement of science and dissemination of advanced
knowledge, but also to encourage the creation of professional
and personal links among the scientists in this international
meeting.
Keeping this objective in mind, we have tried hard - and we
vJi11 try in the next two weeks - to achieve an atmosphere
which will help to attain this goal. certainly we will have
the valuable help of the lecturers, the participants, the LNEC
staff and also of all those very many aspects with which our
country and our people, always surprise those who come from
abroad and wish to know us.
On the behalf of the Organizing Committee I wish to thank
the Lecturers coming from abroad for their efforts and good
will, and the NATO Scientific Affairs Division and the LNEC,
for their logistical support.
To everyone taking part, we offer our best wishes for a
profitable course and a pleasant stay in Portugal.
C HAPTER I
ROCK FILL STRUcrURES: THE PRESENT AN D TH E FUTURE
E. MARANHA das NEVES

From time immemorial man has used rockfill as a construction


material for different structures. For better or for worse,
empiricism always guided design and construction of those
structures; in fact only in the 19505 did all phases of a
construction projection b egi n to be approached in a rational
way.
After the fundamentals of soil mechanics have been
established and some further development in this domain have
been a achieved , attention began to focu s on rockfills. In
this case too, Terzaghi played a leading role when he rightly
disapproved of the construction technique usually adopted at
his time, Le. applying h igh pressure water jets to remove
fines in rockfills for dams construction. A number of
important s teps followed, of which we should mention:

laboratory characterization studies of the mechanical


behaviour of roc kfill, namely their s hear strength and
deformabili ty, which called for design a nd construction
of large-size equipment intended for direct shear
tests, oedometric tests and triaxial shear tests (with
axisymmetry):

acknowledgement of the role of particle fracturing in


the mechanical behaviour of rockfills;

microstructural studies and their important


contribution to sucess of the macrostructural approach:

assessment of the non -li near characteristics of the


Mohr-Coulomb envelope a n d of the possibility of
signi ficant deviation in the normal trend of t h e
initial phase (low stresses) of the stress-strain curve
obtained in 10 compression tests;

importan t results in grain size modelling;

E. Mf" lmha das Nt ...,s (td.). Ad,ym cts;" Rodflll Sm,clltrrs. 1-5.
Cl I991 KII, ....r ' Academic P"b/ishtrs.
2

marked progress in studying rock alteration and


weatherability processes;

increase understanding of the role of water in the


rheologie behaviour of rockfills;

great progress in developing constitutive laws, which,


together with suitable numerical techniques, have
brought about increasing success in description and
prediction of stresses and strains in complex rockfill
structures;
in situ tests for physical and mechanical
characterization of rockfills;
monitoring and observation of the most varied rockfill
structures.
The above steps associated with important developments in
the domaine of quarrying and of transportation, laying,
wetting and compaction of rockfill materials. It is only fair
to emphasize the role played by powerful vibrating rollers in
qualitative advance of rockfill structures.
This progress brought about an impressive worldwide
dissemination of rockfill uses in varied areas such as power
production, agr icul ture, water supply, protection against
natural calamities (earthquakes, slope slidings, floods),
transportation (roads, railways, ports and airports), etc.
What will be the future role of rockfills and rockfill
structures? In an attempt to gather opinions on the matter,
a panel entitled "Rockfill Future" took place during the ASI
on Rockfill structures. Next the main trends of the discussion
are summed up.
First the material itself has to be considered, i. e.
whether it seems there may be large investment in laboratory
tests in future for its physical, mechanical and hydraulic
characterization.
The index properties, for instance, in the near future will
have a growing importance on the estimation of the main
rockfill characteristics.
There is no doubt that the design of rockfill structures,
especially when they are integrated in high precision
projects should always be based on appropriate "design" tests
carried out in situ and in the laboratory and performed on
rockfill samples.
However, in many circumstances, index properties obtained in
stone pieces are of great interest whenever a rough estimate
of the rockfill behaviour is desirable and more sophisticated
tests are not available.
3

Simple structures do not always justify expensive tests and


pre-design analysis of large structures may greatly benefit
from simple tests on small samples that are easy to obtain and
transport, namely access when access to sites is difficult or
outcrops are scarce.
Recent experience shows that the estimate of design
parameters of rockfill is possible although it is not a simple
task. Much more data will be necessary for evaluating the
exact meaning of the first correlations and for assessing the
degree of accuracy that may be expected from this procedure.
Laboratory tests will also have an important role in the
mechanical characterization of rockfills. Certainly we will
not see the design and construction of large apparatus for
testing the real material. As a matter of fact, some
institutions have already done a lot in this very difficult,
expensi ve and complex area to demonstrate that laboratory
mechanical testing of rockfills needs larger equipment than
that used for soils, but the testing of the integral
granulometry is not always necessary. We owe a tribute to
those institutions for this remarkable work.
Laboratory testing in the research area will be mainly
devoted to the study of rockfills consisting of uncommon rocks
(weak rocks included) and of important phenomena such as the
collapse (due to wetting and sUbmersion) and creep of rockfill
as well as its response to cyclic and seismic actions.
Once the question of materials was dealt with, it matters
to discuss what is expected for the different types of
rockfill use, starting by dams.
As a result of recent progress in the understanding of
low-strength rock behaviour, this material will be more widely
applied in dam embankments, which will surely contribute to
more frequent choice of earth-rockfill solutions. Mainly for
high dams, research on the mathematical modelling of the
strain-hardening or strain-softening, elastic-plastic
behaviour, mainly of the core material, as well as on the
f il ter criteria, will make very important contributions to
increase the safety of those structures. The parametric
studies using the f.e.m. will be also an indispensable and
decisive tool.
Concrete face rockfill dams (CFRD), in turn, will attain
heights that had never been reached before. This will be the
result of important progress in the design and construction
technology of the concrete revetments of the upstream slopes,
and avoidance of vertical and horizontal joints. The
recogni tion of the important role of the transition zone
immediately under the concrete slab as a potential filter and
discharge control as well as a support zone with adequate
deformability, will be also an outstanding factor. Moreover
prediction of the movement of the perimetral joint will be
4

improved as a result of progress in the extremely difficult


task of relating dam deformability and the movements in that
joint.
When compared with earth-rockfill dams, CFRD allow a more
versatile use of construction phases for tackling the
important problem of river diversion. It is one more reason
for their increasing use in the future. Their well-known good
behaviour concerning seismic actions is also a feature that
must be taken into account.
One important application of rockfills is on through-flow
or overflow darns. Usually these dams are not very height but
for certain geomorphological and metereological local
conditions their use will be expanded. Knowledge of the true
water pressures in the rockfill medium percolated by the
turbulent through water flow and the mechanisms associated
wi th the scour effect on the structure slopes due to the
overflow are of paramount importance. Theoretical approaches
and physical modelling have been used to tackle those complex
matters but the importance of accumulated experience must be
underlined. wide spread divulgation of results is therefore
asked from those who possess that experience.
Lastly, regarding the employment of rockfills in dams, one
should stress the growing interest of using thin betuminous
cores when the economic factor is determinant. On the other
hand the demand for the use of betuminous revetment in
rockfill dams is not expected to exceed the present level.
To sum up, the future will show an increasing awareness of
the safety of rockfill darns, structures already recognized
among the other possible solutions as the most competitive
from an economic point of view.
An extension of the through-flow structure is the
application of rockfills in spillways as an energy dissipator
element (a solution whose main drawback is the low discharge
capacity). Its use could be advised in some specific
conditions.

Another important area of use for rockfill is related with


the transportation networks. Its use in embankments will
certainly increase owing to the important progress in
construction technologies. Efficient compaction at high rates
of placement when compared with earthfills, (meteorological
condi tions are not an important constraint and embankment
volumes are lower) will have both beneficial economic
consequences and good post-construction behaviour, mainly as
regards creep settlements.
In certain countries there is expected to be an important
application of rockfill in the construction of railways for
high speed trains. In other regions (countries as Australia
and Brazil, for instance) very heavy loads are to be supported
by the railways and special attention must then be given to
the ballast as well as to the embankments. The use of
rockfills as well as the more recent knowledge of their
behaviour will be an important factor in the design of those
structures.

The understanding of the reaction of maritime rockfill


structures to the complex action of the waves will certainly
be an important factor in choosing and improving those
solutions.

Finally as a consequence of research on the rockfill domain


and observation of all types of rockfill structures, steeper
embankment slopes will be employed.
In many cases, waste rockfill type material from mining
works has been accumulated in large areas. As a result of
environmental problems or because different uses are now
envisaged for these zones, special and important issues
related with improvement of those rockfills must be solved in
the future.
Another important feature of rockfill structures (which is
shared with the other fill structures) is their good answer to
the ageing effects (the few exceptions do confirm the rule)
when compared with other types of structures like concrete and
steel, for instance. Because the rehabilitation of aged
structures becomes a very important economic factor nowadays
this is one more reason to choose rockfill solutions.
Prospects about the use of rockfills in structures will
certainly vary according to the region (for instance countries
like U.K. and Germany have practically no dams to be built
which is by no means the case of Australia or Portugal).
Nevertheless, when speaking of all the possible structures all
around the world, there is no doubt that the future will show
us an important increase in demanding for this type of
structures and consequently large efforts will be made
regarding rockfill structures safety (material properties,
laboratory and in situ testing of rockfills, constitutive
laws, mathematical and physical modeling) as well as
construction technology (quarrying, transportation and
compaction of rockfill, use of concrete and betuminuous
material as an watertight element, etc).
For all the presented reasons there is no doubt: rockfills
will have a good future!
CHAPTER 2
PHYSICAL CHARACTERIZATION AND ASSESSr-.1ENT OF ROCK
DURABILITY THROUGH INDEX PROPERTJES
J DHGADO IWOR/CUES

1 - Introduction

The overall performance of rockfill structures is greatly in -


fluenced by the intrinsic properties of the individual rock
fragments. This feature is highly characteristic of rockfills
and, therefore, makes them different from soil structures. In
some specific aspects, it may even be feasible to predict that
behaviour by resorting to simple tests performed directly on
individual rock piec es.
Current practice gives abundant confirmation of this pro-
perty and a good knowledge of the construction material proves
to be very profitable.
Hard rocks make rockfills less deformable and soft rocks are
more susceptible to collapse under wetti ng, and both rock
characteristics are easily assessed by simple and inexpensive
laboratory and field tests.
In some circumstances , rockfills undergo degradation in
time , either in the grain size of rock fragments or in their
strength. In this aspect , the assessment of rock durability is
a fundamental step towards the understanding of the overall
rockfill behaviour.
Among many possible causes, intrinsic rock properties play
decisive roles in this evolution and an adequate understanding
of it should integrate a thorough knowledge of the basic geo-
logical characteristics of rock materials. Bibliography is
full of examples showing that this k nowledge is far from being
of merely speculative or academic interest.
Research on rockfill structures should, therefore, encompass
a certain number of different domains, certainly with close
interrelationships, but using individual methods and knowledge
that may be quite specific to eac h domain.
The inclusion in the research teams of people having back -
grounds in engineering geology, petrography and mineralogy has
this justification and the inclusion of the present paper in
7
E. Murul/ha dos Nl!l'u (~d.), Ad"allces ill Rockfill Slfllc/llreJ, 7_34.
o ]991 KliMa Academic Pllhlishers.
8

this course aims at helping to demonstrate this philosophy. If


any of the data here presented are useful for application to
real cases, the objectives will be completely achieved.

2. Brief presentation of most common rock types

2.1. SIGNIFICANCE OF GEOLOGICAL CLASSIFICATIONS

Geotechnicians know that they have to work with a large diver-


sity of earth materials. Petrography manuals give an almost
endless number of rock names, certainly having real petrogra-
phical meaning, but, most of them, with minor geotechnical
significance when considered on an individual basis.
Grouping materials according to common properties is essen-
tial for communication among different people and classifica-
tions of all types are the most effective tools of this tech-
nical language.
Properties used for grouping are countless and classifica-
tions built with them vary from broad and general to very spe-
cific and partial. When environmental conditions existing at
the time of rock genesis are taken on a rough basis, terms
like plutonic, volcanic, metamorphic and sedimentary are ap-
propriate, although rather generic. However, if silica content
is used in conjunction with them, further subdivisions are
created and increasing precision is introduced. Basic volcanic
and siliceous sedimentary rocks are thus steps towards improv-
ed grouping. Leaving aside other intermediate groups, a large
jump leads to more specific and practical classifications
whose leading properties may be absolutely meaningless in the
upper levels of classification. For geotechnical purposes,
classifications based on the weathering stage are common and
highly relevant, but this parameter is totally dispensable for
classifying the rock as igneous, metamorphic or sedimentary.
Through all this hierarchy of factors, certain inferences
can be made from classifications, at all levels, and even the
highest divisions carry information likely to be useful for
practical purposes, at a much lower level.
Again, igneous is an uninformative term when a geotechnical
design is pursued, but it is enough to add plutonic to allow
much more precise inferences. If this is the case, it would
certainly correspond to a holocrystalline, granular, and, when
unweathered, very resistant and virtually non-porous material.
If made by experienced persons, these inferences can even be
pushed beyond the domain of the material and reach the rock
mass. Aspects concerning the form of occurrence, type of frac-
tures, weathering patterns may be anticipated, and this may
have relevant consequences as regards selection of the most
appropriate methods of field reconnaissance, data interpreta-
tion and, in the end, the design of geotechnical structures.
9

2.2. ROCK MATERIALS IN GENERAL CLASSIFICATIONS

Geology is a scientific domain particularly suited for wide


synthesis and global structuring concepts. Plate tectonics and
geologic cycle are two of them. The former is bringing new in-
sights to the fields of seismology and seismic engineering,
the latter enables us to understand the permanent transforma-
tions affecting our planet.
The first and most general classification of rock materials
draw much of their support from the geologic cycle.
This concept assumes that the Earth is continuously changing
and the existence of a cycle is essential for renovation of
matter over the millions of years of evolution. In general
terms, this cycle is simple and highly explanatory. The lead-
ing idea is that each piece of rock derives from another rock
preceding it and will be transformed into a new one. In con-
crete terms, sedimentary rocks are derived from igneous or
metamorphic rocks (partial cycles may justify one sedimentary
rock being derived from another that is also sedimentary)
through geodynamic processes characteristic of the Earth's
surface (weathering, erosion, transport and sedimentation).
Accumulation of sediments and increasing burial depth lead to
a progressive predominance of geodynamic processes character-
istic of the Earth's internal layers (pressure and temperature
increasing with depth) and to the consequent transformation of
sedimentary rocks into new materials, stable in these environ-
ments. Metamorphic rocks are formed in this way. However, when
pressure and temperature reach high enough values, fusion -
first partial and then total - may take place, and then the
basic requirement for obtaining igneous rocks (the existence
of fused silicate-based material) is attained. From that stage
to solidification, upward migration through the Earth's crust
and from thence to the starting point of the cycle is just a
question of time. But, be it short, as in volcanic rocks, or
long, as in plutonic rocks, time is not a limitative factor
and, in the long history of the evolutive Earth, episodes of
this type are common occurrences and evidence of them are
fully accepted by all Earth scientists.
Low-grade and high-grade metamorphic rocks are transition
terms between metamorphic rocks and, respectively, sedimentary
and igneous rocks. It may be difficult to classify them unam-
biguously in any of these groups, but they are good evidence
of the relationship that exists between all the three groups.
Minerals are the basic solid components of rocks. Most are
crystalline and a few are amorphous. The number of existing
species is extremely high, but those occurring as major rock
components are comparatively very few.
Feldspars and quartz predominate in igneous rocks, and
quartz, carbonates, micas and feldspars predominate in meta-
morphic rocks. Calcite, dolomite, quartz and clays are the
basic constituents found in sedimentary rocks.
10

structures and textures vary widely, and each group has some
very specific features of its own. Some of them may be easily
deduced by any experienced person from this very general clas-
sification.
However, common engineering applications of rock materials
require more precise characteristics, preferably having close
connections with any parameter used in design methods. This
objective leads to the use of geotechnical classifications and
index properties. These items will be dealt with in the fol-
lowing chapters.

3. Properties of rock materials

3.1 GENERAL

Properties of rock materials vary considerably and the causes


of this variation are manifold. To identify the causes and to
understand their roles is very convenient in basic as well as
in practical studies.
First of all, these properties depend on the specific pro-
perties of each specified mineral component. For instance, the
fact of a rock being composed of quartz or calcite will have
quite distinct implications as regards its properties.
But differences in texture and structure also have great
influence, even within materials having identical mineral com-
positions. Differences in mineral grain size, in their habit
or in their arrangement (foliated, with random distribution,
etc.) are some of the ways in which minerals influence rock
properties.
Besides the dependence on the mineral matter, rock proper-
ties also depend on the empty part of the rock. Voids are re-
levant entities which, in some aspects, may even be more de-
terminant than the solid components.
Texture and structure of void space are also influencing
factors in rock behaviour. Voids may be more or less equidi-
mensional (pores) or very flat entities (fissures) and the
predominance of one of these types has far-reaching implica-
tions as regards rock properties. Plutonic and metamorphic
rocks are fissured materials, and sedimentary rocks may be
predominantly porous. Fissures reduce strength and ultrasonic
velocities, and increase permeability much more efficiently
than pores.
In clastic rocks, cementing matter plays a decisive role in
rock properties and frequently has more influence on mechani-
cal properties than the clastic components. Silica, calcite
and clays are common binding materials in clastic sedimentary
rocks and the predominance of any of them determines the rock
properties from the very beginning of deposition, through dia-
genesis and up to the metamorphism environments. Besides the
II

direct influence on strength, they are decisive factors in


rock durability.

3.2. ROCK MASSES AND ROCK MATERIALS


Another point that should be taken into account is the distin-
ction between rock masses and rock materials, namely as re-
gards their specific relevancy in the different phases of
study and design of any geotechnical undertaking and particu-
larly of rockfill structures.
Rock mass is applied to in situ bodies of rock and it in-
cludes the rock, its network of discontinuities and its varia-
tions on the state of weathering. Unless the rock be very weak
(either because of its lithological nature or its state of
weathering), most relevant properties of rock masses, to a
great extent, depend on the actual fracture network. Strength,
deformability and permeability are strongly correlated with
discontinuities and may be largely unrelated to the properties
of the rock material itself.
For any work to be carried out in situ (dam foundation,
shaft, quarry), rock mass is the geotechnical concept that
should be the leading idea underlining the studies of explora-
tion, in situ characterization and geotechnical design. In the
context of the present paper, this concept will not be pushed
too far, since only the extracted materials will be dealt
with, in spite of it being recognized that studies of the
foundation of rockfill structures and of quarry selection are
basically concerned with rock masses, and thus strictly con-
nected with the purpose of this document.
The raw material for construction of any rockfill structure
does not conserve large discontinuities, such as faults and
joints, unless they are tightly cemented, and only the discon-
tinuities (usually fissures and small fractures) of a high
pervasive character are expected to accompany the rock from
the quarry to the rockf i 11 . In this case, discontinuities
become an integral part of the rock material and should be
considered as one of the rock components.

3.3. BRIEF CONSIDERATIONS ABOUT SAMPLING

Representative samples of rock masses cannot be extracted and


only in situ tests can aim at characterizing these large scale
entities. Samples can hardly represent small parts of large
discontinuities but may adequately represent the rock material
with its network of small-scale fissures and fractures. This
important distinction has relevant consequences on the defini-
tion of suitable tests and makes feasible the laboratory test-
ing of rockfill materials.
12

Al though feasible, representative sampling is far from being


an easy task to accomplish, and it needs a thorough reflection
by the person concerned about definition of the sampling plan
before starting to collect ad hoc, eventually meaningless,
samples. Accuracy of test results reflect the care put into
sampling and the study reports should always comment on the
real significance of the samples used.
Sedimentary sequences frequently have layers with different
lithologic composition and physical properties. A thin clay
layer may eventually have great influence on the stability of
a quarry slope but it may be almost meaningless as regards
influence on the properties of the quarried material. The pur-
pose of the study and the actual site conditions determine
what and how to sample a specific lithologic entity in any
given circumstances.
On the other hand, igneous and some metamorphic sequences
may present slight to large variations in weathering stages at
a scale of the quarry face. Although the rock characteristics
in a given area may present variations compatible with the re-
quirements of the slope design and extraction methods, they
may be too variable if they are to be used as construction ma-
terial in certain exigent works. Protection rockfills may be
one of these uses.
Results of laboratory tests are frequently used to charac-
terize a whole quarry production, but without always assuming
that this way consists of extrapolating to populations far
larger than the tested samples.
This limitation may not be possible to overcome completely
with larger samples, and the number of tests cannot be in-
creased beyond reasonable limits. This means that accuracy
will always remain a questionable parameter in any plan of
rock materials characterization.
However, it should be stressed that simple and quick tests
may be used to complement those plans because they allow the
identification of eventual existing sub-populations, guide
sampling for more demanding tests and turn easier their sub-
sequent interpretation.
Porosity, quick absorption ratio and point load strength
come immediately to our mind, but visual inspection by anyone
experienced in petrography and geology may be far more infor-
mative than any of the other tests mentioned. This allows the
construction of a "picture" of the entire area which will act
as the basic framework for the extrapolation operations with
the laboratory tests.
site investigation, sampling operations and interpretation
work are three components that should be combined by the dif-
ferent intervening specialists, always bearing in mind that
this may be more easily done when a geological and petrograph-
ical basis is used as a bridge between the different tasks.
13

4. Laboratory characterization of rock materials

4.1. GENERAL

Laboratory characterization of rock materials may be achieved


through an almost endless number of tests, running from the
very simple and fast to the very sophisticated and time con-
suming.
The significance of test results varies according to the
type of rock concerned and their relevance from a geotechnical
viewpoint changes according to the objective, namely the use
that is expected for the rock material. This means that it is
not the number of tests that is meaningful but the criterion
of choice that can point to the quality and relevance of any
laboratory testing programme.
The philosophy used for grouping and presenting rock pro-
perties depends on the objective and may vary widely.
The Commission on standardization of Laboratory and Field
Tests of the International Society of Rock Mechanics (ISRM)
used the following way of grouping tests (ISRM, 1979):

category I: Classification and characterization


Category II: Engineering design tests

Density, porosity, point load strength, permeability and


sound velocity are examples of Category I properties, while
triaxial and direct shear behaviour and time-dependent and
plastic properties are grouped as engineering design proper-
ties.
The RILEM Commission 25-PEM, "Protection et Erosion des
Monuments", has followed a different approach and has used the
following groups (RILEM, 1980):

I. Tests defining the rock structure (v.g. porosity and perme-


ability) ,
II.Tests defining the properties connected with the presence
and movement of water (v.g. saturation coefficient and water
absorption by capillarity),
III. Tests defining the internal cohesion (v.g. dynamic modu-
lus, tensile and compressive strength),
IV - Tests to determine the mechanical surface properties
(v.g. hardness),
V. Durability tests (v.g. salt crystallization and frost re-
sistance) ,
VI. Miscellaneous (v.g. thermal expansion).

In rockfill structures, rock elements interact with each


other and some shape and size properties currently used for
characterizing aggregates for road construction and for con-
crete may be fully justified. other appropriate classifica-
tions could, thus, be mentioned here.
14

The scope of the present paper does not aim at an exhaustive


treatment of the aspects concerned with rock characterization
but rather at presenting a comprehensive approach to what
might be considered a "feasible recipe for characterizing rock
materials for rockfill purposes".
with this in mind, a specific grouping of rock properties
has been devised for this paper. It includes the following
items:

- intrinsic properties
- index properties
- shape and size properties
- durability properties

Typical engineering design properties will be dealt with in


another paper and will not be considered here.

4.2. INTRINSIC PROPERTIES

Intrinsic properties are the "identity card" of any material


concerned. They should include a name and some basic and very
specific characteristics that may be considered as "all-purpo-
se" elements, irrespective of the use expected for that mate-
rial.
The petrological name and description should always accompa-
ny any study of rock materials. It may be obtained by current
petrography means existing in most laboratories for testing
materials. Petrography data are very informative and specifi-
cations of materials to be used for aggregates and other pur-
poses currently recommend this study. Many organizations, na-
mely ASTM, BS and ISRM, have standards with guidelines for pe-
trographic examination of rocks.
Petrography can hardly give enough information on the porous
space, a fundamental rock component. Porosity, a simple and
inexpensi ve parameter should therefore integrate the "identity
card". Specific gravity is also recommended.
The above considerations by no means assume that those pro-
perties take constant or rigid values. Mineralogic composition
may vary within certain limits for a given quarry or even for
a sedimentary layer, and porosity may vary from sample to sam-
ple and even from piece to piece. The choice of these proper-
ties, rather than for their rigid values, is justified for
their simplicity and informative character and assumes that no
other index and design tests should be made or interpreted
without having information on those intrinsic properties.
Fissuration patterns, pore size distribution and intrinsic
permeabili ty are very meaningful intrinsic properties but
their experimental determination is not available in all test-
ing laboratories. Although they are intrinsic, they may not be
included on the "identity card".
15

4.3. INDEX PROPERTIES

4.3.1. General. A large number of rock characteristics may be


grouped as index properties. The main idea underlying this
concept is that these properties characterise the response of
the rock specimens to a given external stimulus; they are cur-
rently used for classification but are not directly used for
design calculation purposes. Point load strength, hardness and
aggregate impact value are examples of index properties.
In this definition, the use to be given to the property con-
cerned is relevant, meaning that a certain property may be in-
cluded in more than one group. For instance, porosity may be
taken as an "intrinsic" and "index" property, and compressive
strength as an "index" and "design" property.
Index properties are easy to ascertain and are available in
most laboratories for testing materials. It is their simplic-
ity and the possibility of achieving large number of results
that makes them so attractive for researchers and practitio-
ners concerned with rockfill structures.
For the purpose of the present paper, only rockfill struc-
tures are considered and this fact reduces the number of index
properties that should be taken as relevant and might be pre-
sented in great detail.
Some properties also carry information useful for durability
assessment and can thus be grouped under such an item, but no
single property is sufficient basis for a jUdicious durability
assessment. These properties will be dealt with in a specific
section later on in this paper.

4.3.2. Some common index properties Index properties are


currently determined by means of simple and easy to carry out
tests and are primarily used for classifying rock materials.
Although in different ways and to a varying extent, they carry
much potential information and when properly used enable an
inference analysis to be made for obtaining meaningful ele-
ments about other more complex or difficult-to-obtain proper-
ties. It is in this precise context that they act as indices
of something placed at a higher level of difficulty or com-
plexity and it is in this context that their name is justi-
fied.
Testing methods will not be presented in detail except when
considered indispensable for the current discussion. When ap-
propriate, quotations will be used to refer to the adequate
source.
Porosity is a fundamental parameter of rock materials. It
is well defined and may be determined with high accuracy by
using common laboratory methods (ISRM 1979; RILEM 1980).
Real density or grain density characterizes the solid matter
of the rock and together with porosity they give a good pictu-
re of the material.
16

Bulk density, maximum water content and void ratio can be


computed by using equations of interdependence with porosity
and real density and in this way they do not bring additional
information but only a different way of presenting what is al-
ready known.
water content after a given immersion time under normal con-
ditions of P and T gives a rough picture of the rock and, when
properly calibrated, it may be a very fast way of obtaining
elements of high practical interest. In his pioneer work,
Hamrol (1961) used a two-hour immersion water content for
characterizing the weathering stage of some Portuguese gra-
nites, and correlations with some mechanical tests fully bore
out that method. When used wi thin a gi ven rock type, this
water content presents high correlations with maximum water
content and therefore with porosity. Fig. 1 gives one example
illustrating this type of correlation.
In sedimentary rocks, porosity shows close correlations with

W max ('II.)
6

/
5

3
/
/
2 /
1 ~
/
/
o
o 1 2 3 456
W 2h ('Yo)

Figure 1 - Relationship between water content at 2h and


at saturation in granitic specimens

other properties, as has been frequently shown (v.g. Mamillan


(1958), Tourenq et Archirnbaud (1974), Smorodinov et al.
(1970. Fig. 2 shows an example taken from carbonate rocks
illustrating some details that may accompany such correla-
tions.
Correlations between porosity and ultimate compressive
strength show two different branches. One applied to specimens
17

with porosity higher than about 3 to 4 percent, while a second


branch embraces those having lower porosity.

'0
PO"<OSI r y ('/.)

Figure 2 Relationship between porosity and compressive


strength for dried specimens of limestones

High porosity carbonate rocks commonly correspond to mate-


rials having more or less equidimensional pores while those of
low porosity are predominantly fissured materials. Fig. 2
clearly illustrates that fissures are much more efficient in
reducing strength than a similar volume of pores. That figure
shows that it may also be expected that a carbonate rock with
less than 3 percent porosity and less than 130 MPa of compres-
sive strength is certainly a fissured material, otherwise its
strength would be much higher.
Velocity of transmission of elastic waves, i.e. sonic velo-
city, can be obtained non-destructively and is a very common
laboratory method for characterizing rock materials. Some la-
boratories have transducers with pointed end-pieces which
allow them to be used on rough surfaces, namely in the field.
Sonic velocity is an elastic property of solid materials and
can be theoretically computed for mul tiphase materials by
weighing the sonic velocity of each component with its res-
pective percentage of occurrence. This theoretical sonic velo-
city is always higher than the experimental values because
real materials are not "perfect" and have fissures and pores
that transmit the elastic waves far more slowly than the solid
matter does.
within the same rock type, sonic velocity presents good cor-
relations with several other parameters owing to its sensiti-
vity to the presence of voids. Many authors have made use of
this parameter, either as an objective in itself or as a means
to obtain information about other more complex parameters.
18

Tourenq et Fourmaintraux (1974) use sonic velocity to com-


pute indices of quality and claim to be able to distinguish
the amount of porosity of the fissure-type (nf) from that of
the pore-type (np) and have designed some abacus appropriate
to this objective.
Following similar principles, Delgado Rodrigues (1982) de-
fined an index of fissuration (IF) specially adapted to fis-
sure-prone rocks (marbles, granites and other crystalline
rocks). It uses the sonic velocity of the "ideal" rock without
voids, computed with the modal composition, and the experimen-
tal velocity in dry and saturated conditions. As might be ex-
pected, IF and nf present very high linear correlation (LNEC
1987) .
Examples of the use of sonic velocity for obtaining estimat-
ed values of engineering design parameters are also frequent
in the literature.
Several laboratory tests have been designed for obtaining
information on the strength properties of rock materials. The
preference for one or another is usually based on the avail-
ability of testing facilities and on the type of samples
available for testing.
Compressive strength is the best known, but it requires
fairly large samples for extracting regular specimens and
presents a high cost for the preparation of specimens for
testing.
Crushing strength is used by many laboratories for testing
rock materials for rockfill structures. It requires a platen
press and approximately spherical rock lumps. Preparation of
specimens is the most time consuming step. Some classifica-
tions of materials for rockfill purposes use this parameter
for grouping (Marsal and Resendiz, 1975).
Point load strength can be easily determined in the field
and in the laboratory and increasingly has supporters of its
use as an index property. It may be obtained in irregular spe-
cimens, although regular ones give more accurate and reproduc-
ible results. ISRM (1985), has a recommended procedure for ob-
taining it and some other researchers have suggested slightly
different ways of testing. For instance, Guifu and Hong (1986)
suggest the use of the area of the rupture plane to compute
the point load strength instead of De 2 (De = equivalent dia-
meter) as recommended by the ISRM standard. The physical mean-
ing of using the area of the rupture plane is straightforward
and, although most times the results obtained by both ways are
correlated, that fact may justify the preference for direct
determination of the rupture area.
Point load strength (PLS), correlates pretty well with com-
pressive

strength (0 r ) but anisotropy may introduce

some dis-
tortlons ln the most commonly suggested correlatlon formulae.
Frequently, the formula

Or = 25 PLS
19

is referred to, but results as low as

Or = 13 PLS
were found in the course of a study of greywackes (Delgado Ro-
drigues and Jeremias, 1989). Difficulty in having fully repre-
sentative samples motivated by the occurrence of lithologic
anisotropy and fissure and weathering heterogeneities account
for this abnormal correlation factor. Water content influences

both r and
PLS and should,

therefore, be taken into account
when testlng rock materlals.
Basically, PLS and crushing strength depend on the same rock
property - its tensile strength - and eventual preferences for
one of them mainly rely on the merit of the specific testing
characteristics. Simplicity, portability and versatility are
good reasons to prefer the point load test.
Some very well known tests are addressed to the characteri-
zation of the properties of rock surfaces and they may be in-
cluded in the generic term of hardness tests.
Schmidt hammer value (rebound hardness) is obtained by a
very simple and quick procedure, but it is rather imprecise
and highly dependent on the surface finishing state. The re-
quired equipment is very light and portable, but values should
be taken as rather rough approximations. Sawn and polished
surfaces substantially increase the reliability of test re-
sults.
Los Angeles abrasion value (abrasion hardness) is the weight
loss of a rock sample in the Los Angeles abrasion machine in
specified test conditions. It is a fairly reliable value, but
it is sensitive to the size of tested pieces. A very wide ex-
perience of using this parameter exists and many users con-
tinue to accept this test in spite of the large dimensions of
the samples required. Because each sample consists on a large
number of rock lumps, its preparation may take into account
the diversity of existing rock varieties (grain size, weather-
ing stage, lithologic variations, etc.) so that it is an easy
way of averaging the results and reducing the range of possi-
ble values. Obviously, when this is the objective, sample pre-
paration has to be made judiciously and the fact must be
pointed out in the test report.
Slake durability test was primarily designed for predicting
rock performance, in particular low strength types (Franklin
and Chandra, 1972). However, its simplicity makes it attrac-
tive for laboratory testing even outside that initial domain
of application. Its application to stronger materials requires
an increase in the testing duration, but it may prove to be
very interesting as a controlling parameter for changes in
stone resistance (Delgado Rodrigues, 1976). A few other expe-
riments have also shown that it may come into use as an index
test, namely for study of rockfill materials (Delgado Rodri-
gues et al. 1982, Veiga Pinto et al. 1986).
20

In another category of tests, those resorting to dynamic


actions on particulate materials may be grouped. The aggregate
crushing value and the aggregate impact value are examples of
these tests (Collins and Fox, Ed. 1985). They are quite common
in the characterization of aggregates for concrete and for
road construction, but they are much less popular in rockfill
studies.

4.3.3. Estimation of rockfill characteristics through index


properties. The design of rockfill structures, especially when
they are integrated in exigent undertakings, should always be
supported on appropriated "design" tests carried out in situ
and in the laboratory and performed on "rockfill samples".
However, in many circumstances, for instance during the
feasibility stage, index properties obtained on stone pieces
are of great interest whenever a rough estimation of the rock-
fill behaviour is desirable and no more sophisticated tests
are available.
Simple structures not always can justify the execution of
expensive tests and pre-design analysis of large structures
may greatly take benefit from simple tests carried out on
small samples easy to obtain and transport, namely when the
accessibility of sites is poor or the outcrops are scarce.
From the LNEC experience, estimation of design parameters
of rockfills is possible although not being a simple and
straightforward task. Much more data is necessary for evaluat-
ing the exact meaning of the first correlations and for as-
sessing the degree of accuracy that may be expected from this
procedure.
The first attempts to find useful correlations were made
with a strength parameter (crushing strength) and a hardness
value (slake test), in the assumption that both bulk strength
and surface properties of the stone pieces are important for
the overall behaviour of the rockfill although their exact in-
dividual roles might not be easily ascertained.

strength and deformability properties of rockfills, even of


laboratory constructed samples, depend on a large number of
factors which may be grouped in intrinsic and extrinsic. Rock
properties (e. g., strength, porosity, mineralogy, surface
hardness), shape and size of stone pieces are intrinsic fac-
tors whi Ie the placement conditions, compaction degree and
confining pressure are extrinsic. water content, an external
and variable factor, is another relevant condition and should
therefore be taken into account when these correlations are
sought for.
This large number of factors cannot be considered simulta-
neously and some of them should be kept constant. In the first
attempts, only the intrinsic properties were considered and
within this group it was assumed that shape and size were also
approximately constant. Placement conditions were similar for
21

70
Weathered schist and graywacke 10 (d",)
0(') (saturated)

0 1 6 Id",)
Sound graywacke .. (saturated)
60
6 Sound basak o (d",)

Sound granite Id",)


50 o Dolostone II (dry)

40

10 15 20
Pa (kN)

Figure 3 - Correlation between crushing strength (Pa) and the


angle of internal friction ()

all tests and samples were tested at the maximum compaction


stage. Figs. 3 to 5 assemble the results taken from several
LNEC pUblications (Delgado Rodrigues et al. 1982, Veiga Pinto
et al. 1986, LNEC 1987, 1989,) and they illustrate the at-
tempts so far experienced.

400
rn
tL
Weathered schist and greywacke 1
0 Id"ll
(saturated)

~ 61'''11
Sound graywacke 1 (saturated)
"0 300
'a"
lJ.J Scorlaceous basalt

Sound granite

200
Do Iaslane 1 .. (dry)
x (saturated)

100
o


10 15
Pa (kN)

Figure 4 - Correlation between crushing strength pa) and oedo-


metric modulus (Eoed)

Crushing strength was the first strength index to be used


but it is progressively being SUbstituted by point load
strength (Quinta Ferreira et al. 1990). Correlations here
presented still use the first parameter.
22

70
0 Id~J
0(") Weatl1ered schist and greywacke / (saturated)
[J
6. (dry)
60 Sound graywacke / A (saturated)

" Sound basa!t 0 (dry)

50
... Sound graMs _ ~dry)

0 Dolostone II (dry)

40

30
00 05 10 15

Slake (2nd cycle) (% loss)

Figure 5 - Correlation between slake durability value (2nd


cycle) and the angle of internal friction ()

In a broad approach, strength of stone pieces shows fair


correlations with the strength parameter (phi) and with de-
formability (oedometric modulus) of the tested rockfill sam-
ples.
Correlation between slake values and the strength parameter
(phi) show a clear trend but some points that fall outside
this trend give a certain amount of uncertainty to this corre-
lation. New data are required before definite conclusions can
be drawn and it should be admitted that new data may come to
demonstrate that splitting data in homogeneous groups defined
by their petrographic parenthood might be a required way to
improve the accuracy of estimations.
Slake test, in particular when hard rocks are concerned,
may be influenced by subtle differences in the sample surface
since the addition or the subtraction of one more "grain" to
the small amount of the lost residue may be enough to intro-
duce large differences in the final results. Longer duration
of abrasion cycles reduce the relative errors and this may
also be a necessary refinement to be introduced in this test-
ing method.

4.4. SHAPE AND SIZE PROPERTIES

Rockfills are particulate bodies whose constitutive elements


may vary within a very large size range. The same descriptive
concepts used for fine soils are applicable to rockfills but
their experimental materialization is highly cumbersome with
large dimension pieces. Particle size curves, mean diameter
uniformity coefficient, etc. are useful concepts and the over-
all behaviour of rockfill structures depend very much on the
23

actual values of these parameters. Other papers will discuss


this matter in more detail.
Shape of quarried materials is highly dependent on the geo-
logical and petrographic conditions of the quarry site, name-
ly, on the attitude and spacing of joints and bedding planes.
Massive and isotropic rock bodies (granites and other plutonic
rocks, some limestones, etc.) tend to give equidimensional
(cuboidal) blocks, while anisotropic rock bodies (gneiss,
greywackes, schists, etc.) may tend to supply flaky and elon-
gated particles.
Shape indices of individual particles may be computed by
measuring the three orthogonal dimensional axes of each parti-
cle or be qualitatively evaluated by comparing with typical
forms of appropriate shape-classifying abacus.
Roundness is not a relevant parameter in quarried materials,
but may be significant in sedimentary gravel deposits. Quar-
ried materials are always angular, while gravel deposits may
display round to very well rounded elements.
Size, shape and roundness may play appreciable roles when
the hydraulic behaviour is concerned. These concepts will be
more thoroughly dealt with in another contribution to this
course.

4.5. DURABILITY

4.5.1. General remarks. All rock materials decay at a geologi-


cal time scale but some of them degrade much faster and may
display visible effects of degradation after very short de-
lays. This degradation may induce relevant perishing of their
expected properties and consequently affect their performance
as construction materials.
To be able to forecast how any given rock material will be-
have when integrated in any engineering work is a long-stand-
ing aim of researchers.
The ways that have been followed so far are multiple and
diverse as a direct consequence of the high complexity of the
subject, the disparity of researchers backgrounds and inter-
ests, the variability of rock materials and diversity of the
utilizations concerned.
Requirements regarding rock stability are different accord-
ing to the use expected for the material. Stability is much
more important in protection rockfills than for rockfill
structures, but it is even more decisive in ornamental rocks
and in artistic artifacts.
When dealing with geotechnical works, namely with rockfills,
the ambition ascribed to this type of forecast is usually sa-
tisfied with qualitative judgements and with some simple clas-
sifications of the expected behaviour (v.g. "poor", "good",
"excellent"). However, it should be strongly emphasized that
such types of qualitative judgement need to be supported in
quantitative testing data together with qualitative analysis
24

and require an adequate background of experience with similar


situations.
In the scope of the present paper, weathering and weather-
ability of rocks in the geological context will not be dealt
with, but it should be pointed out that events occurring dur-
ing the geological history of any material may have trans-
formed it to an extent that may be very decisive afterwards
when conditions of the rock environment are modified, in par-
ticular when materials are exposed or quarried.
without being exhaustive, a brief analysis of the leading
ideas on rock durability assessment may be of some interest to
readers.
The soundness test (salt crystallization test) and freeze-
thaw test are well known ways of predicting the performance of
rock materials. However, in spite of their interest, it seems
that their utilizations have been pushed far beyond their li-
mits of validity. In particular, for most circumstances that
use these methods for prediction, a more or less important
lack of parallelism between the decay mechanisms acting "in
situ" and in the laboratory test may exist and this fact may
seriously impair the validity of the results for forecasting
the real behaviour of materials. Significant differences in
test results may arise from slight changes in testing condi-
tions namely in drying operations, for the salt crystalliza-
tion test (price 1978) and in the rate of temperature de-
crease, for the freeze-thaw test (Lautridou et Ozouf, 1978).
Simulation tests and accelerated ageing tests are current
procedures for assessing rock durability and are both support-
ed by the aggravation (slightly, in simulation and drastical-
ly, in accelerated tests) of the ambient conditions. It is
assumed that the aggravation does not significantly change the
decay process, but in some tests this may prove to be a rough
assumption.
Existence of real similarity between the decay processes
acting "in situ" and in the corresponding laboratory test has
always been a strong reason for accepting the results of dura-
bility assessment studies. And, very likely, this works quite
well in extreme sides of the quality ranking, i.e., in good
and bad classes of materials. However, when intermediate qual-
ity rocks are to be evaluated, even slight errors may prove to
be technically and economically very meaningful.

4.5.2. Some methods of rock durability assessment. Test re-


sults may be used in manifold ways, but durability assessment
always needs to be supported by the knowledge of real work
behaviour, either gathered directly by the technician con-
cerned or borrowed from the literature. As time goes by and
experience increases, new ways of assessing durability appear,
aiming at obtaining better correlations between predicted and
observed behaviours.
25

Some of those ways rely on the similarity existing between


the mechanisms acting in tests and in the real situation, as
is the case, for instance, of frost resistance assessment.
These tests might be described as "performance tests".
Other ways are preferentially used for characterization of
some properties deemed to be informative as regards durabili-
ty. The basic assumption of this way of durability assessment
is that with a good knowledge of the rock properties and an
adequate level of background experience it is possible to as-
cribe the concerned material to specific groups whose future
behaviour can reasonably be predicted. The relevant data used
for this purpose are index properties and all the quantitative
results are indirect information on the expected rock durabil-
ity performance. This way of rock durability assessment will
be dealt with in more detail in the next section of this pa-
per.
This very simple division into two typical ways of tackling
durability assessment may bring relevant practical consequen-
ces, namely by introducing the necessity of clarifying which
option is the one to be followed in any specific test.
This system is based on the philosophy of approaching the
problem and not necessarily on the specific tests used for
that purpose. As an example, the Los Angeles wearing test was
designed for assessing the performance of rock materials in
road and railroad construction and a certain similarity can be
pointed out as existing between the test and reality: abrasion
is a real decay process in ballast and pavement base as well
as in the test. However, when used in concrete, rocks are not
subject to abrasion wear, but the Los Angeles test is recom-
mended by most common standards concerning concrete. There is
nothing wrong with this use but it is important to clarify
that in these circumstances the Los Angeles abrasion loss is
used as an index property and, therefore, only as indirect
information on the stone durability.
Literature makes many references to tentative use of metho-
dologies of durability assessment, some of which will be enu-
merated subsequently.
Farran et Thenoz (1965) suggest the use of permeability,
specific surface and the water-rock reactivity for prediction
of "in situ" performances of rock masses. Irrespective of the
intrinsic meaning and validity, these methods are far from
easy to carry out and, therefore, they are more research ori-
ented. struillou (1969) recommends the use of hydrogen peroxi-
de envisaging the assessment of the role of clay minerals.
with a similar purpose but using microscopy methods, were
the ways followed by Scott (1965) and weinert (1964, 1968).
Franklin and Chandra (1972) and Gamble (1971) have made wide
use of the slake durability test and Delgado Rodrigues (1976)
has adapted this same test to hard rocks by adding a pretreat-
ment with ethylene-glycol and increasing the duration of the
test.
26

Except the microscopy methods, the tests referred to above


introduce a more or less aggressive treatment in order to
evaluate its incidence on the material concerned. These treat-
ments do not necessarily aim at replicating any real decay
mechanisms and the criteria used for ranking the tested rocks
rely on experience and use the results obtained with rock ma-
terials of known real behaviour (good and bad, when possible)
to compare with the results under appreciation.

4.5.3. Assessment of rock durability through index properties.


In order to be of practical interest, the assessment of rock
durability should be feasible in a short operation and as far
as possible be supported on simple and easy to carry out
tests. The best suited methodologies use index properties to
reach this desideratum.
Swelling strain in conjunction with ultimate strength was
suggested by Olivier (1976) for classifying the Karroo mud-
rocks and Felix (1987) made use of swelling strain in conjunc-
tion with the ratio between ultimate strength in wet and dry
conditions for assessing the durability of Swiss molassic
sandstones. These methods attribute a leading role to clay
minerals and the proposed index properties aim at quantifying
it. The methylene blue value is also directed to evaluation of
the amount of clay minerals and several authors have been
using it for assessing rock durability (Ngoc Lan, 1980; Stapel
and Verhoef, 1989).
Other authors have proposed more complex indices, two of
which may be found in smith et ale (1970) and Fookes et ale
(1988) .

These indices are, respectively:

01
DAR
1 + w

where,

DAR durability absorption ratio


01 durability index (wet abrasion test)
W water absorption

and,

I S (50) - 0.1 (SST - 5WA)

where,
27

static rock durability indicator


average compressive strength in dry and saturated
conditions
SST weight loss in the soundness test with magnesium
sulphate
WA water absorption (according to BS 812)
SGsSd saturated bulk density (according to BS 812)

In our laboratory, we have been for sometime trying to


define appropriate ways of having quick answers of practical
value to questions concerning assessment of rock durability.
In some cases, owing to difficulties concerning the existing
methodologies of assessment, and in others owing to limita-
tions in the available testing facilities, we have been led to
develop some very simple alternative ways of tackling this
problem. When first trying to systematize the situation re-
ferring to carbonate rocks an abacus for the geotechnical
classification was developed (Delgado Rodrigues, 1988). It has
porosity and swelling strain as the only parameters and some
applications so far available support the validity of this
procedure (Fig. 6). Both parameters are accessible to most
laboratories' testing facilities and the equipment required is
simple and inexpensive .
.---
V AV
20
Porosity (%)

IV AIV BIV CIV DIV

10

III Alii Bill CIII

5 DI
II All BII
2.5 CI
AI BI
2.5 5 10 20
A B C o
Swelling strain ,010. 4

Figure 6 - Abacus for the geotechnical classification of


carbonate rocks (from Delgado Rodrigues 1988)
28

These parameters evaluate two distinct rock properties, the


amount of pore space and the importance of clay minerals pres-
ent in the rock. They react mainly to two distinct decay mech-
anisms, which are salt crystallization and changes in water
content. The roles of these parameters, though clearly dis-
tinct, are not mutually exclusive, but rather a certain syner-
gy may exist when both parameters increase simultaneously.
Engineering works often deal with carbonate rocks, namely
when selecting construction materials in sedimentary regions.
The selection of the most appropriate sites for borrow
quarries is often one of the first objectives in design, a
task that is easily accomplished with this simple classifica-
tion.
The classes defined in the proposed abacus show good correl-
ation with actual mechanical properties, but it is the clear
representation of the real behaviours of some known carbonate
stones that is most striking. Fig. 7 shows one example of
classification of some of the tested samples.

Porosity (%)

IV @

10~--+---+-------~--------------~

III

5r---r---+-------~

II
2.5~

2.5 5 10 20
A B c D

Swelling strain 10. 4


KEY:
1 . "L ioz" limestone
2 . "Coimbra" dolostone
3 . "An~a" limestone
4 - "Keddara" glauconitic limestone
5 - "Mondego" argillaceous limestones and marls

Fig.7 - Example of the utilization of the classification


abacus (from Delgado Rodrigues 1988)

Worth mentioning are the differences between the "Lioz"


limestone, probably the best Portuguese limestone, the "Anc;:a"
29

limestone, a very fast decaying stone when placed in contact


with saline solutions, and "Mondego" argillaceous limestones
and marls, used in some protection rockfills with poor perfor-
mance. "Lioz" and "Anc;:a" are clay-free stones and, in spite of
the great difference in porosities, their very low swelling
strain is, consequently, very low, as might be expected. "Mon-
dego" stones are argillaceous, their swelling strain is high
and their "in situ" behaviour is in perfect agreement with the
DIV class in which they are included.
The experience gained with this classification leads us to
conclude that class A corresponds to clay-free materials,
whereas in C and D classes, clays tend to be the decisive rock
consti tuents. Class B may be considered as a transitional
group.
Although simple, this classification scheme may be helpful
in definition of the first requirements to be fulfilled in
purchase and construction contracts. In rockfill construc-
tions, for instance, the different classification fields can
be used for selecting the different stones for each specific
place in the embankment. For protective rockfills, a possible
specification document might stipulate that only materials
from AI and All categories would be acceptable, and that a
certain amount (to be specified) of transitional BI and BII
materials could be admitted for that purpose.
In the follow up of our research on durability assessment
of rock materials, another attempt was made, trying to intro-
duce a certain degree of universality by constructing one
index that could be applied to any type of rock without the
need for auxiliary abacus.
The underlying assumption of the new index is that swelling
strain and porosity are the basic parameters for expressing
the importance of the most harmful components of the rock ma-
terial. This is true in many (or even most) common situations,
but this assumption requires a critical analysis whenever any
specific practical application is to be made.
It was considered that together with those two parameters
the inclusion of a strength parameter might be helpful, on the
assumption that two rocks with similar swelling strain and
porosity would behave differently should they have different
strengths.
with this in mind the basic formula of that index (Delgado
Rodrigues and Jeremias, 1990) is:

IRD
n + 2a

where,
30

IRD index of rock durability


R ultimate compressive strength
Rt strength parameter for comparative purposes
n porosity (%)
a the mantissa of the swelling strain when expressed as:

For keeping the index adimensional, Rt = 1 MPa. If different


units are used for expressing strength, Rt should be converted
according to those units.
In order that the index might keep its comparative value,
the swelling strain only can be expressed as mentioned, which
means that the value used corresponds to the actual swelling
strain multiplied by 10 4 In this way, for most common rocks,
the weights that correspond to swelling and porosity in the
proposed formula fall within the same range (between 0 and
30), and therefore, variations in each parameter may notice-
ably influence the final value of the index.
From the above mentioned paper, it can be seen that the
first trials with this index are very promising. When studying
the greywackes from the upstream slope protection riprap of
Beliche dam, it was found that samples with IRD lower than 4
correspond to stones with poor behaviour in accelerated ageing
tests and should therefore be discarded for such utilization.
The shortage in the available data prevented a more thorough
discrimination in that classification.
The application to the limestones dealt with in the other
approach (Delgado Rodrigues, 1988) allows us to formulate the
following conclusions:
- one of the best Portuguese limestones (the "Lioz" limestone)
has an IRD of about 94, while some porous clayey specimens
with extremely poor behaviour have IRD lower than 1, going
down to 0.2;
- "Anc;:a" limestone, very porous, though clay-free, decays very
rapidly, in particular when in contact with salt solutions.
This rock is very well known from some Portuguese masterpiece
monuments, particularly for its softness and high decay rates.
The corresponding IRD value is about 1;
- common Portuguese micritic limestones, known to be good con-
struction materials, have IRD above 5, and most of them above
15.
In a first approach, for carbonate rocks, it was considered
that values lower than 2 signify low durability, while values
greater than 10 correspond to materials having good to excel-
lent durability. A larger number of results is necessary in
order to clarify the transition zone between 2 and 10.
31

Acknowledgements
Some of the presented results were obtained by collaborative
work with Maranha das Neves, veiga pinto, M. Quinta Ferreira
and F. Telmo Jeremias.

References
COLLIS, L. and FOX, R.A. (Editors), 1985 - Aggregates: sand
gravel and crushed rock aggregates for construction pur-
poses. Published by the Geological Society, London, 220p.

DELGADO RODRIGUES, J., 1976 - Estimation of the content of


clay minerals and its significance in stone decay. Proc. 2nd
Int. Symp. on Deterioration of Building Stones, Athens,
pp.105-108.
DELGADO RODRIGUES, J., 1982 - Laboratory study of thermally
fissured rocks. Proc. 4th Int. Congo on Deterioration and
Preservation of Stone objects; Louisville (Ky.), pp. 281-
294.

DELGADO RODRIGUES, J., 1988 - Proposed geotechnical classi-


fication of carbonate rocks based on Portuguese and Algerian
examples. Engineering Geology, Vol. 25, No 1, pp. 33-43.

DELGADO RODRIGUES, J. and JEREMIAS, F.T., 1989 - Character-


ization of greywacke-like rocks as construction materials
(in Portuguese). Proc. 3 2 Encontro Nacional de Geotecnia,
Porto, pp. Al19-130.

DELGADO RODRIGUES, J. and JEREMIAS, F.T., 1990 - Assessment


of rock durability through index properties. Proc. 6th Int.
Congo of the IAEG, Amsterdam, pp. 3055-3060.

DELGADO RODRIGUES, J., VEIGA PINTO, A. and MARANHA DAS NE-


VES, E., 1982 - Rock index properties for prediction of
rockfill behaviour. Proc. 4th Int. Congo of the IAEG, New
Delhi, pp. VI 39-47.

FARRAN, J. & THENOZ, B., 1965 - L'alterabilite des roches,


ses facteurs, sa prevision. Annales de l'ITBTP, XVIII eme
annee, Nov. 1965.

FELIX, C., 1987 - Essais et criteres de choix pour des gres


(molasses) de SUbstitution lors de travaux de restauration.
Chantiers / Suisse, 18, (5), pp. 419-423.
FERREIRA, M. Q.; RODRIGUES, J. D.; PINTO, A. V. and JEREMIAS,
F. T., 1990 - Evaluation of strength in irregular rock lumps
32

for characterization of rockfills. Proc. 6th Int. Congo of


the IAEG, Amsterdam, pp. 3119-3124.

FOOKES, P.G., GOURLEY, C.S. and OIKERE, C., 1988 - Rock


weathering in engineering time. Quat. Jour. of Engineering
Geology, 21, pp. 33-57.

FRANKLIN, J.A. and CHANDRA, R., 1972 - The slake-durability


test. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., vol.
9, pp. 325-341.

GAMBLE, J.C., 1971 Durability-plasticity classification


of shales and other argillaceous rocks. PhD Thesis, Univ.
of Illinois.

GUIFU, X. and HONG, L., 1986 - On the statistical analysis


of data and strength expression in the rock point load
tests. Proc. 5th Int. Congo of the IAEG, Buenos Aires, pp.
833-894.

HAMROL, F.A., 1961 - A quantitative classification of the


weathering and weatherability of rocks. Proc. 5th Int. Conf.
Soil Mech. Found. Engng., Paris, pp. 269-274.

ISRM, 1979 - suggested methods for determining water content,


porosity, density, absorption and related properties and
swelling and slake-durability index properties Int. J. Rock
Mech. Min. Sci. & Geomech. Abstr., vol. 16, pp. 141-156.

RILEM, 1980 - Recommended tests to measure the deterioration


of stone and assess the effectiveness of treatment methods.
Materiaux et Constructions. Vol. 13, No 75, pp. 175-253.

ISRM, 1985 Suggested method for determining point load


strength. Int. J. Rock. Mech. Min. Sci. & Geomech. Abstr.,
vol. 22, No 2, pp. 51-60.

LAUTRIDOU, J.P. and OZOUF, S.C., 1978 - Relations entre la


gelivite et les proprietes physiques (porosite, ascension
capillaire) des roches calcaires. Proc. Int. Syrup. on De-
terioration and Protection of Stone Monuments, UNESCO/RILEM,
Paris, Pap. 3.3.

LNEC, 1987 - Paradela dam. Study of the rockfill and struc-


tural behaviour analysis (in Portuguese). Internal report,
230/87 - NF/NP-DG, LNEC, Lisboa.
LNEC, 1989 - Barragem da Apartadura. Aterros experimentais e
ensaios laboratoriais (in Portuguese). Internal report,
99/89 - NF-DG, LNEC, Lisboa.
33

MAMILLAN, M., 1958 - Methode de classification des pierres


calcaires. Annales de l'ITBTP, n Q 125, Paris, Mai, 1958.

MARSAL, R. and RESENDIZ, D., 1975 - Presas de tierra y enro-


camiento. Pub. Editorial Limusa, Mexico, D.F., 546 p.
NGOC LAN, T., 1980 - L'essai au bleu de methylene, un pro-
gres dans la mesure et Ie controle de la proprete des
granulats. Bull. Liaison des Lab. Ponts et Chaussees, 107,
pp . 13 0 -13 5 .
PRICE, C.A., 1978 - The use of the sodium sulphate crystal-
lization test for determining the weathering resistance of
untreated stone. Proc. Int. Symp. on Deterioration and Pro-
tection of Stone Monuments. UNESCOjRILEM, Paris. Pap. 3-6.
OLIVIER, H.J., 1976 - Importance of rock durability in the
engineering classification of Karroo rock masses for tunnel-
ing. Proc. Symp. on Exploration for Rock Engineering,
Johannesburg. Edited by Z. T. Bieniawski, vol. 1, pp. 137-
144.
SCOTT, L., 1955 - Secondary minerals in rock as a cause of
pavement and base failure. Highway Res. Board, Proc.,
1955, pp. 412--417.

SMITH, T., McCAULEY, M.L. & MEARNS, R.W., 1970 - Evaluation


of rock slope protection material. Highway Res. Board (323),
Nat. Res. Council, US Nat. Ac. of Science.

SMORODINOV, M.I.: MOTOVILOV, E.A. and VOLKOV, V.A., 1970 -


Determinations of correlation relationships between strength
and some physical characteristics of rocks. Proc. 2nd Congo
ISRM, Beograd.

STAPEL, E.E. & VERHOEF, P.N.W., 1989 - The use of methylene


blue adsorption test in assessing the quality of basaltic
tuff rock aggregate. Engineering Geology, vol. 26, pp. 233-
246.

STRUILLOU, R., 1969 - Prevision de l'alterabilite des mate-


riaux en fonction de leurs caracteristiques propres et de
leurs utilisations. Colloque de Geotechnique, Toulouse,
Mars, 1969.

TOURENQ, C. et ARCHIMBAUD, C., 1974 - Proprietes des calcai-


res. Proc. 2nd Int. Congo of the IAEG. Sao Paulo (Brasil)
Pap. IV-19.
34

TOURENQ, C. et FOURMAINTRAUX, D., 1974 - L'indice de quali-


te des roches, quelques applications. 2nd Int. Congo of the
IAEG, Sao Paulo (Brazil), Pap. IV-20.

VEIGA PINTO, A., DELGADO RODRIGUES, J. and MARANHA DAS NE-


VES, E., 1986 - Some improvements in the characterization
of basalts and limestones for rockfill purposes. Proc. Vth
Congo of the IAEG, Buenos Aires, pp. 1469-1475.

WEINERT, H., 1964 - Basic igneous rocks in road foundations.


CSIR Research Report, 218, Bull. Nat. Inst. Road Res.,
pretoria, 5.

WEINERT, H., 1968 - Engineering geology for roads in South


Africa. Engineering Geology, 2 (6), pp. 363-395.
CHAPTE R 3
ROCK FILL MODELLING
A. K. PARKIN

I. Introd uction

Rockfill may be defined as a coarse-grained and frec-draining material won by


quarrying rock. Therefore its mOSt obv ious features are (or should be) the coarse,
angular particles and the absence of pore-pressure.
The coarseness of the part icles and their interlocked state makes il impossible to
sample and test rockfill as for other soils, so that laboratory test specimens must
always be scaled down to some degree, as well as being Te-const ituted by compaction.
The test samples must also be of small size, in relat ion to prototype, and affected by
boundary and seating conditions. For these reasons, the exten t to wh ich laboratory
teSlS can be relied upon to predict field behaviour has a lways been a rather open
questio n and probably remains so . Therefore, it is the purpose of this paper to look
at some principles that should guide a laboratory investigation in order to elim inate
these uncertainties as far as possible.
One such set of principles for preparing a laboratory model rockfil l ha s been
formulated by Fumigalli et al. (1970 ) as follows:

(a) Equality of strength in all fractions


(b) Geometrically simi lar grading curves
(c) Equality of void ratios
(d) Equality of Old r.lIios (defined later)
(e) Similarity of panicle shape

Whil st all these condilion s are logicliL nOI all are capable of being mel in ull C:lses,
nor has it been proven thilt this is necessary. In fact, it is clear that some measure of
latitude exists (Charles and Wail s. 1980).
In their consideration of modelling requi rements, Marachi et al. (1972) noted. on
the basis of Henzian theory (as d id Lowe. 1964). that the frict io nal strength and
compressibi lity of an assembly of particles should be quite independent of panicle
size. In respect of model gradings. they wen I on to note earl y applications of scalping

"
E. Maranlw das N~"t!$ /~d. ). Ad\'QI1("~s in RocJ:fil/ S,mC',ures. 35-51.
e 1991 KllOwr AC'u"~mic P"t>lis"~rs.
36

on Goschenenalp Dam, Switzerland (Zeller and Wulliman, 1957) and scaling on


Shihmen Dam, Taiwan, (Lowe, 1964).

2. Rockfill in Place

In the era of dumped rockfills, there was no restriction on the size of spalls that could
be accepted, and indeed large material was most desirable. With the move into
compacted rockfills, material must be spread in layers of 1 m or less for effective
compaction, which limits the maximum spall size to something rather less than this
(e.g. 2/3; Penman and Mitchell, 1970), oversize material being removed by raking.
More often, however, rockfills will be smaller than this, with maximum particle sizes
of 150 mm or so, particularly if derived from tunnel spoil. In all cases, however, these
sizes are well above the capability of normal laboratory testing equipment, so that the
determination of strength and deformational parameters must come from either the
substitution of a model rockfill of reduced size, or the use of large test equipment, or,
as proves to be the only realistic course, a combination of both.

3. Triaxial Testing Equipment

The first use of large scale triaxial equipment for testing rockfill appears to have been
by Hall for the Isabella Dam project (US Corps of Engineers, L.A.), as reported by
Leps (1970). This, and other test facilities operating in North America, are tabulated
by Leps, with sample diameters of 0.2 up to 1.13 m and stone sizes generally limited
to one quarter of thisl. To this list, some others can be added, as in Table 1.
From these lists, it might be concluded that the majority of rockfill testing is done with
300 to 500 mm samples, with equipment larger than this either becoming extremely
expensive, or being restricted to vacuum loading.

4. Maximum Particle Size, Test Samples

The maximum particle size of a model rockfill, d, is normally determi~ed by the


smallest dimension of the test sample, D. Thus, in terms of the Did ratIo, Penman
1971) states that the lower limit is 4 for a broad grading (possibly on the basis of data
assembled by Leps, 1970), or 6 for a narrow grading (as used by Marsal, 1973l The

1. The 1.13m and 0.9m machines at CFE Mexico and UC Berkeley, California, are no longer used
for rockfill testing. Rocklill testing at CFE is now based on O.3m diameter samples.

2. Fumigalli et aI., (1970) advise a value of 10, although current ISMES practice uses a lower limit
of 5. Marachi et a!., (1972) use 6 for aggregates with C = 5.
ll
37

basis for such figures, although rarely discussed, probably lies in the work of Holtz and
Gibbs (1956) who were able to obtain a consistent Mohr envelope for Did ~ 6, but
obtained higher values of f for Did = 4.

Table 1

Location Sample dia Max Pressure


mm kPa

Monash University, Australia 380 1500


CSIRO Div. of Geomechanics
Melbourne, Australia 380 1500
Snowy Mountains Engineering
Corporation, Cooma Australia 570 1700
Norwegian Geotechnical
Institute Oslo, Norway 625 <100
Imperial College London, U.K. 300 900
Building Research Station U.K. 230 700
ISMES, Bergamo, Italy 350 & 500 2000 & 1000
LNEC, Lisbon, Portugal 305 2000
Inst. f Bodenmechanik u
Felsmechanik, Karlsruhe, FRG. 1000 2200
AIT, Bangkok, Thailand 250

The only other laboratory study that has set out to study Did as it affects <1>' is that
of Lewis (1956), testing loose single-sized crushed granites in two shear boxes (60 and
300 mm). These results show <1>' deviating from a common trend at a value of Did =
40 (Fig. 1). Whilst conditions within the rigid confines of a shear box are undoubtedly
more severe than for a triaxial test, or even oedeometer compression, this figure is
clearly excessive for use in triaxial modelling. It is also clear that empirically-based
values of Did, as above, have been producing plausible results for some years.
The choice of Did is probably determined to a greater degree by the effect it has on
compaction for the production of a sample of representative density and void ratio.
Oversized material leads to restricted mobility within the sample and is likely to result
in large voids at the sample perimeter (if not elsewhere). This may lead to membrane
indentation errors, or even perforation, as wel1 as causing density to be low. Clearly,
the oversize will become less critical with more finer material to flow into these voids,
or in the case of soft rocks where the necessary fines may be created during
compaction. On such a basis, Marachi et aI., (1969), as quoted by Charles and Watts
(1980), justify their selection of Did = 6, provided that not more than 30% of the
particles are in the maximum size range, essential1y conforming to the earlier work of
Holtz and Gibbs (1956).
38

50
Triaxial tests (av.>
45

--
300 x 300 Shear box
0'
40

35
Limit for
large box
30
01 10 10 100
Particle Size- mm

Fig. 1. Direct shear tests on rockfill (Lewis, 1956)

Apart from the Did issue, there is also the question of the absolute effect of
maximum particle size on <1>'. This receives rather more attention in literature, but
remains inconclusive, probably due to the difficulty of isolating the relevant parameters
(Charles and Watts, 1980). On theoretical grounds, Marachi et aI., (1972) argue that
friction parameters should not be particle size dependent, but on the basis of nine
investigations reviewed, there is a clear trend for <p' to decrease with increasing d.
Their own results later confirmed this (Fig. 2).

o 55
OROVILLE DAM MATERIAL
&
.
~ r---
D/d=6
-...
c:
.2

--
u
I.J...
~ ~

------
o 45
...c: 2900

-
~

--
c:
..... 4490
a "...
.!!
g' 35
<l: 10 25 50 100 250
Maximum Particle Size

Fig. 2. Variation of <P' with d in triaxial testing (after Marachi et aI., 1972)
39

The only investigators who have found <1>' to increase with d are Lewis (1956), who
seemed surprised that it should be so (Fig. 1), and Charles and Watts (1980) who
recorded a small increase in <1>'. From the variations shown in Fig. 2, it might be
concluded that d should not be too much less than 100 mm in order to be within a
degree or so of the field value of <1>'.

5. Model Gradings

Once the maximum allowable particle size (d) has been determined, a test mix
grading can be established in either of two ways (Fig. 3):

(i) by parallel scaling


(ii) by scalping
Parallel scaling may be the logical choice and is often advocated (eg. Fumigalli,
1969). In this case the field grading curve is translated to the left by a constant factor,
such that the maximum particle size is reduced to d. This is most likely to be
accomplished by screening out component sizes and re-combining to form the scaled
grading, whereupon care has to be exercised to ensure that these finer components are
still representative of the main rockfill in terms of hardness, strength and shape
(Fumigalli et aI., (1970). This is not necessarily the case as these fine components
often form from inferior material.
It may also be found that a scaled grading may contain a disproportionate quantity
of fines, which, apart from their possible effect on <1>', may cause the rockfill sample
to be no longer free-draining, a matter of quite serious consequence if large triaxial
specimens are being used.

100~----------,---------~~----~-----'

Field
grading

oL---~----~~----------~~-----------
1 10 100 1000
Particle Size (mm)

Fig. 3 Grading modification


40

Alternatively, the model grading may be produced by scalping (eg. Zeller et aI.,
1957; Fagnoul, 1969), wherein all material coarser than d is discarded. The percentage
passing any size smaller than d is then scaled up by a constant factor. It is just as
necessary as before to ensure that this scalped sample is representative of the field
rockfill, and if this should be in doubt then the laboratory sample should be prepared
by crushing down large stones. It might also be necessary to scalp the lower end of
a scaled grading if problems are encountered with drainage (as above) .
. The scaling choices will depend on the type of rock and its field grading. For hard
rock and narrow gradings, such as the crushed basalts of Marachi et aI., (1969), parallel
scaling is appropriate. For broader gradings with more fines, scalping may be more
suitable to avoid the problem of excessive fines, whilst for soft rocks which degrade
during compaction the scaling method may become unimportant. Most rocks will in
fact degrade to some degree on compaction such that precise control of model grading
is not possible, but, as noted by Charles and Watts (1980), there is no evidence that
such variations significantly affect <1>'.
The reliability of such modelling procedures must be evaluated by comparison with
field performance. This could be done by back-figuring from a slope failure, but slope
failures in rockfill are almost unknown and backfiguring is not without complications.
However, a very useful evaluation is possible from the work of Valstad and Str~m
(1976), who compared their laboratory data with a series of in situ plate load tests, in
connection with Svartevann Dam investigations. These plate load tests (analysed by
the Brinch-Hansen formula) indicated a value of <1> (for heavily rolled granitic gneiss)
of around 53, which proved to be substantially in excess of laboratory triaxial results
of around 45. It may still be arguable which figure is more credible.

6. Some Particular Issues relating to the Oedometer Test

6.1 Oedometer Dimensions

Because rockfill is a highly frictional material, substantial shear stresses will develop
on the wall of an oedeometer during a compression test, so that the applied platen load
is not transmitted equally to all parts of the sample. It is possible for as much as 50%
of the applied platen load to be lost into the wall in this way. One possible means of
overcoming this problem is by the use of a compressible ring vessel, built up from
alternating rings of steel and a compressible material (such as cork or rubber), as in the
ISMES consolidometers of Fumigalli (1969). These machines, 600 and 1300mm dia.
with HID = 2 (and another 500 dia. at LNEC), rely on an initial calibration in
compression, which may not necessarily remain valid with a compacted sample in
place. Nor is there a description in literature of the type of support used during sample
formation. Although evidently successful, this concept has not been adopted elsewhere.
41

Rigid-walled oedometers, on which most rockfill compressibility testing is based,


are characterised by their HID ratio and whether or not the ring is fixed or floating
with respect to the base platen (in addition to the internal diameter, which determines
the size of stone that can be accommodated)3. Wall friction development will be
minimised by the use of a floating ring arrangement (Fig. 6), or in some cases by the
use of a lubricant, such as Marsal (1973), who noted that axial strains could increase
by 20% with lubrication in a fixed ring test. However, in spite of the considerable
problems of alleviating frictional resistance, rockfill in situ generally proves to be
significantly stiffer than predicted on the basis of oedometer tests.
With regard to the HID ratio, there are conflicting requirements such that a
compromise must be established. Clearly, this ratio should be kept small (say 0.5) to
minimise wall friction in relation to the total axial load. On the other hand, seating
conditions (against the top platen in particular) will generally be less than ideal, and
their influence on the load-deformation relationship will be minimised by keeping HID
large (1 or greater). For this reason, an HID ratio of 0.25 as in standard geotechnical
laboratory work would be unacceptable, whereas HID = 2, as used by Frassoni et al.
(1982), should largely eliminate seating errors.

Vertical stress MN 1m2


o 1 2

2
0~

c:

.......
0 4
III

"0
....u... 6
~
8

10

Fig. 4. Stress-strain curves from oedometer tests

3. Frassoni et al. (1982) state that it is ISMES practice to limit stone size to D/IO, and show from
comparative tests that D < 500mm may lead to some error.
42

On completion of compaction, the top surface of a specimen will be quite uneven


and of doubtful unifonnity. This will therefore require trimming, which is liable to
cause some disturbance to adjacent rockfill, followed by capping, often with sand. The
platens may also include filter elements for use in making penneability measurements.
These factors together can lead to a situation where the capped region is considerably
softer than the very stiff rockfill elsewhere, which may then cause significant errors in
the early portion of the load-defonnation curve. Whether this is a factor in the
considerable variations of shape that are observed in laboratory load-defonnation
curves (as in Fig. 4) cannot be detennined as capping arrangements are rarely
described in literature. However, some indication can be obtained from the curves of
Fig. 5, obtained on replicate samples of Hornfels, with cappings as indicated. On this
basis it would seem prudent to use a generous value of H, or a capping of maximum
stiffness, such as cement or plaster, to minimise possible errors from this source.

Vertical Stress (MPa)


0 1 2

Lysterfield Hornfels
-~ p=22 tlm3 H=610mm

E
E
2
-:~~ Floating ring test

c I
.Q
+--
0 4
E...
.....0Cll
0
6 Sand
capping
_L
I

Fig. 5. Effect of capping on oedometer response

6.2 Side Friction Models

The result of an oedometer test is nonnally presented as a stress-strain curve wherein


the vertical stress <Jv is computed at mid-height from the applied load plus the weight
of platen and fittings and the top half of the sample. This, the nominal vertical stress,
will be in error by the amount of load shed in side friction, which causes the true
43

vertical stress to be both complex and non-unifonn. However, a representative average


vertical stress can be derived by either of two procedures, one in which side friction
is estimated from a final jacking test (Penman and Charles, 1976), and the other
wherein skin friction is established from replicate fixed and floating ring tests (Parkin
and Adikari, 1981).
In the first of these methods, it is considered that the average side friction stress
(presumably on upper and lower halves) in a floating ring test is given by
(1)

where K" is the coefficient of earth pressure at rest and 11 the coefficient of friction
of rockfill on the oedometer ring, the product 11K" lying within a narrow range (0.06
to 0.2) for most situations4 For substitution in Eq. (1) average stresses are computed
as

(2)

(3)

where L is the applied axial load and S is obtained from the final ring jacking test
(evidently as 2S). Substituting also d = 2h for the BRE oedometer and assuming S/2
to be small with respect to L leads to
(4)

and this inserted into Eq. (2) gives a value of cry corrected for side friction:

a (
- L 1 - -IlKo) /-d
1t 2 (5)
" 2 4

On the basis of a final jacking test on a sandstone, Penman and Charles (1976) found
that ilK" = (l.ll). indicating a side friction correction of ).7%.
Whilst thi, figure is entirely plausible, it cannot be readily verified as laboratory
predictions rarely match field peIi'ormance with this precision. Several other issues
must also be u\dressed, among them the fact that the jacking test, presumably
perfoIllled \\ ith the sample uncler axial compression, cannot be associated with any
particular v.tiue of L. and will have a distribution of shear stress that is very different
from the flo.pi"b ring test. Furthermore, the value of K for a compacted sample is not

4. These figures arc evidently drawn principally from Kjaemsli and Sande (1963), who record wall
loads in fixed ring test at 12 to 40% (average 20%) of axial load.
44

constant, but decreases with load and only approaches Ko at high stress. Finally,
whereas K can be validly applied to average stresses, a friction coefficient cannot, as
indicated in the fact that Eq. (J) cannot apply near the sample centre, where 't~0.
The alternative procedure was developed by Parkin and Adikari (1981), based on
an oedometer de~igned to operate in both fixed and floating ring modes, with provision
for the measurement of total wall load (F) in the latter cases. For each mode, a
distrihution of shear stress was asslImed to reflect the probable shear displacements,
as in Fig. 6.

Fixed Ring Floating Ring

Fig. 6 Oedometer side-friction model

In terms of the measured value of F, the corrected value of crv at mid depth will be

ov - o>,(lIOm) - a FIA (6)

where A is the sample area, and <X = 0.75 and 0.25 for the fixed and floating rings.
This, however, is not the average stress on the samples, and the computation of average
stress should use <X = 0.67 and 0.17 respectively. If these corrections are applied to
the nominal stress-strain curves for the fixed and floating ring tests, a reasonable co-
ordination should result. The effectiveness of this may be judged from Fig. 7, noting
that further improvements might come from other assumed distributions of wall shear
stress. 6
5. The N(d ocdomclcr is also designed 10 opera Ie in both fixed and floating ring modes, and on the
basis of all avcr;\ge side friclion of 20% in Ibe fixed mode, a correction of 10% has been applied
10 !ltC Il(lal illg rillg mode (Kjaemsli and Sande, 1963). At Karlsruhe (Table I), the load transmitted
to the toltol1l platcn is recorded and used to evaluate side friction.

6. In practice, platen displacements are invariably not equal in a floating ring test (base being half or
less top), suggesting that the "neutral axis" may not be at mid hcight and that a value of ex = 0.17
may be low. This may result from greater density towards the sample basco
45

15.-------------------------------~~.,
o Fixed ring
o Flooting ring
a..
~

<f)

-
<f)

~
Ul
'0 05
u
:;:
'-
Q)
>

o 02 OA 0'6 08 10
Vertical Strain %

Fig. 7. Corrected stress-strain curves, Talbingo Rhyolite


(Parkin and Adikari, 1981)

6.3 Effects of Initial Stress

The production of samples for laboratory compressibility testing normally involves


rockfill being placed in a rigid containment vessel and compacted in place with a
pneumatic hammer. At the end of this process, the vertical stress can be presumed to
be at, or near, zero, but the locked-in horizontal stress is substantial. This condition
will not necessarily reproduce the in situ stress state for rockfill compacted by rolling,

2
~
Sugarloaf Mudstone
....
<I>
p= 21 tlm 3
:l
If)
If) Fixed ring test
....
<I>

-
a...
.c
....
0
w
......
0
....:
......
<I>
0
U

a 05 10 15
(Jv MPa

Fig. 8. Coefficient of earth pressure


46

but to secure measurements for such a comparison is not an easy task.7


If O"h is required from an oedometer test, the most usual way is to strain gauge the
containment ring. However, to obtain reasonable sensitivity, the ring must be fairly
thin (6mm in the case of the Monash Oedometer), whilst retaining sufficient rigidity
to resist distortion during compaction. Strain gauges will normally be placed at several
levels, such as quarter points, and used to evaluate an average lateral stress. If
monitored by level, then a vertical distribution of O"h may be derived, but the accuracy
of this will be questionable because of the bending stiffness of the rings. It is also not
easy to calibrate the strain gauges as fitted, although the manufacturers calibration is
probably adequate.
By relating average lateral pressure to mid-height vertical pressure, the development
of K in a stage load test can be traced as in Fig. 8 (with the hatched area giving an
indication of local variability). At low stress, K is clearly determined by initial
compaction conditions, but evidently converges on a true value of K" for stresses of
sufficient magnitude (in this case - 0.5 MPa). For this test HID = 1 and wall friction
was measured at 27% of axial load for the purpose of axial load correction
If settlement observations are plotted in the form of an e-Iog 0" relationship, it will
be observed that the curve shapes are very similar to those normally obtained on clays.
This similarity is such as to suggest that the traditional Casagrande construction might
be applied to determine a "pre-consolidation pressure", although the point so located
could not have the normal meaning. From Fig. 9, it can be observed that the point of
maximum curvature (marked) clearly increases with increasing density, and because
this density is a function of the compactive effort imposed it must be concluded that
the marked points are varying with initial lateral stress. It can be further noted that
this transitional behaviour occurs at a stress similar to that at which the K value
stabilises on Fig. 8.
These relationships show that some aspects ofrockfill performance in the oedometer
are affected by initial compaction stresses, and if these are not reproducing stresses in
situ, then some care may be required in interpretation.

7. A Case History

Blue Rock Dam is a central core earth and rockfill structure on the Tanjil River in
south-east Victoria, Australia. At the damsite, the valley is incised into Lower
Devonian mudstones and sandstones, overlaid at higher levels by weathered Tertiary
basalt, both strata being utilised for rockfill production (Murley, 1981).
7. Whilst most laboratories use air hammer compaction, the Karlsruhe facility uses static compaction,
which results in some fairly massive lateral pressures. This clearly creates a different type of pore
structure, at comparable global void ratios, and the procedure that is most relevant to practice is
a matter of conjecture.

8. This is not a problem for ISMES consolidometeT. Here approximately one ring in seven is
straingauged, and the measured values of crh appear to be good insofar as they correlate well with
triaxial tests via the Jaky formula (Frassoni et a!. 1982).
47

Load (kPa)
50 100 200 500 1000 2000
-40

'38
Sample No.3
pd = 200tlm 3
(I) 36
0
+=
0
a: 34
'0
0 Sample No.4
> '32 pd - 210 t/m 3

'30

28

Fig. 9. Void ratio - pressure relationships for mudstone, Blue Rock Dam

During field rolling trials, it was judged that the mudstone was the superior material,
although subject to thermal breakdown. It was designated as select rockfill and placed
in the upstream shoulder (3A). The basalt, being of low grade, was designated as
random roekfill and placed in the downstream shoulder (3B). The rolling trials
indicated that a density of 2.2 tlm3 could be achieved from four passes of a lOt
vibrating roller.
For triaxial testing, both materials were supplied sieved out into component
fractions, which were re-combined to prescribed gradings that were in principle scaled,
but modified somewhat to reduce an anticipated problem with fines. For these test
mixes <\nax = 75 mm, giving DId = 5, but there remained some 30 to 40% finer than
4.7 mm, extending down to dust (Cu around 1(0). During mixing and placement, it
was found to be difficult to prevent segregation, the effects of which are evident in Fig.
lOa. However, by increasing the number of layers from 5 to 8, a much improved
sample resulted (Fig. lOb). It will be clear from such photographs why the laboratory
density of 2.1 tlm3 is likely to be less than that from field rolling, although the internal
structure may still be representative.
For the first of the samples tested (Fig 6a) drainage proved to be so bad that loading
had to be extended over a period of four weeks. Hence for subsequent samples of the
3A roekfill, all material passing 4.7 mm was eliminated (and that passing 1.2 mm for
3B roekfill), by means of which testing was possible within a day.
However, whereas a density of 2.1 tlm3 was achieved with reasonable effort in the
3A material, it was found to require extreme effort in the case of the 3B material.
48

Fig. lO. Rockfill triaxial specimens, Blue Rock Dam (Parkin, 1983)

Subsequently, on testing, the 3B material, whilst considered to be the poorer of the


two, gave a slightly higher strength (Fig. 11). This problem appears to have arisen
because one of the fractions of the 3B basalt was vesicular, causing the particle
packing to be much tighter at the specified density, and indicating that the smaller
fractions are not always representative of in situ material.

8. Conclusions

Since serious testing of rockfill began around 1960, it has been necessary to reach a
compromise between available size of test equipment and the tolerable degree of
particle scaling. Whilst in theory any level of particle scaling should be of no effect,
this cannot be relied upon and most people prefer not to scale too far.
Model gradings will generally be derived on the basis of scalping or scaling, or
some other compromise which may be dictated by the special circumstances of a
particular rockfill. Whether this is representative of rockfill in place may depend more
on the compactive effort or density than on particular details of grading, particularly
as these will normally be altered during compaction. Maximum particle size is limited
by the chosen DId ratio (4 minimum, but preferably 6), which is determined by
practical requirements for the production of a uniform representative sample.
49

0
2
- - 3A Rockfill
- - 38 Rockfill
-------,'-
a.. "'-
:?:
If)
""- \
If)
\
....
Q)

en \
.... \
0
Q)
..c
(f)

0 2 3 4 5
Normal stress M Pa

Fig. 11. Mohr circles for triaxial tests, Blue Rock Dam.

In the oedometer compression test, modelling also involves the design of the
oedometer and its principal dimensions, and the provision made for minimising and/or
correcting side friction. It is also emphasised that care should be taken to avoid
introducing a soft end condition that could distort observed stress-strain curves, and
further noted that some aspects of behaviour will be affected by the initial stress state.
In summary, the evidence would seem to indicate that predictions of <1>' from large
triaxial tests are mostly believable, although in the absence of a slope failure (which
even then is not easily backfigured) there is no real check. By contrast, settlement
predictions from oedometer tests are easily and almost invariably checked against field
performance, and found in most cases to under-predict rockfill stiffness in situ, despite
the considerable efforts made to minimise side friction.

Acknowledgement

Data on rockfill testing for Blue Rock Dam is included with the permission of the
Rural Water Commission of Victoria (Australia), for which the Author expresses his
thanks. In addition, Figs 5 and 8 are drawn from student projects by H. K. Lim and
Y. C. Wong and by G. Hesketh.

References

Charles, lA. and Watts, K.S. (1980). Influence of confining pressure on shear strength
of compacted rockfill. Geotechnique 30:4 pp. 353-367.
Fagnoul, A. (1969). Shear strength. 7th Int. Conf. Soil Mech. and Found. Engg.
Mexico. Spec. Session No. 13, pp. 191-200.
50

Frassoni, A., Hegg, U. and Rossi, P.P. (1982). Large-scale laboratory tests for the
mechanical characterization of granular materials for embankment dams.
Proceedings, 14th Int. Congr. on Large Dams, Rio de Janeiro, 4 : 727-751.
Fumigalli, E. (1969). Tests on cohesionless materials for rockfill dams. Proceedings,
ASCE, v. 95 no. SM1, Paper 6353, pp. 313-330.
Fumigalli, E., Mosconi, B. and Rossi, P.P. (1970). Laboratory tests on material and
static models for rockfill dams. Proceedings 10th Int. Congr. on Large Dams,
Montreal, 1:531-551.
Holtz, W. G. and Gipps, H.J. (1956). Triaxial tests on pervious gravelly soils.
Proceedings, ASCE, v.82 no. SM1, Paper 867.
Kjaernsli, B. and Sande, A. (1963). Compressibility of some coarse-grained materials.
Proceedings, European Conf. Soil Mech. and Found. Engg. Wiesbaden, 1 : 245-251.
Leps, T.M. (1970). Review of shearing strength of rockfill. Proceedings, ASCE, v.96
no. SM4, Paper 7394, pp. 1159-1170.
Lewis, J .G. (1956). Shear strength of rockfill. Proceedings, 2nd Australia - New
Zealand Conf. on Soil Mech. and Found. Engg. pp. 50-52.
Lowe, J. (1964). Shear strength of coarse embankment dam materials. Proceedings,
8th Int. Congr. Large Dams, Edinburgh, pp. 745-761.
Marachi, N.D., Chan, CK., Seed, H.B. and Duncan, J.M. (1969). Strength and
deformation characteristics of rockfill materials. Report TE-69-5, Univ. of
California, Berkeley.
Marachi, N.D., Chan, CK. and Seed, H.B. (1972). Evaluation of properties of rockfill
materials. Proceedings, ASCE, v.98 no. SM1, Paper 8672, pp. 95-114.
Marachi, N.D., Seed, H.B. and Chan, CK. (1969). Strength characteristics of rockfill
materials. 7th Int. Conf. Soil Mech. and Found. Engg., Mexico, Spec. Session No.
13, pp. 217-224.
Marsal, R.J. (1973). Mechanical properties of rockfill. In: Embankment-Dam
Engineering, Casagrande Volume (R.C Hirschfeld and S. J. Poulos, Eds.), pp 109-
200.
Murley, K. (1981). Blue Rock Dam Project, Tanjil River, Victoria. AN COLD
Bulletin No. 58 (February), pp 55-61.
Parkin, A.K. and Adikari, G.S.N. (1981). Rockfill deformation from large-scale tests.
Proceedings, 10th Int. Conf. Soil Mech. and Found. Engg. Stockholm, 4:727-731.
Parkin, A.K. (1983). Strength and compressibility of rockfill. Blue Rock Dam Project,
Tanjil River. Report to State Rivers and Water Supply Commission of Victoria, No.
83/1, Dept. of Civil Eng., Monash University.
Penman, A.D.M. and Mitchell, P.B. (1970). Initial behaviour of Scammonden Dam.
Proceedings, 10th Int. Congo on Large Dams, Montreal, 1: 169-187.
Penman, A.D.M. (1971). RockfiIl. Building Research Station Current Paper 15nl
(April), lOpp.
Penman, A. and Charles, J.A. (1976). The quality and suitability of rockfill used in
dam construction. Proceedings, 12th Int. Congr. on Large Dams, Mexico City, 1 :
533-566.
51

Valstad, T. and Str0m, E. (1976). Investigations of the mechanical properties of


rockfill for the Svartevann Dam. Norwegian Geotechnical Institute Pub!. No. 110,
pp 3-8.
Zeller, J. and Wulliman, R. (1957). The shear strength of the shell materials for the
Goschenenalp Dam, Switzerland. Proceedings, 4th Int. Conf. Soil Mech. and Found.
Engg., London 2:399-404.
CHAPTER 4
LABORATORY SHEAR STRENGTH TESTS AND THE STABILITY OF
ROCKFILL SLOPES
1. A. CHARLES

1. INTRODUCTION
The laboratory measurement of the shear strength of recompacted samples
of rockfill is reviewed with emphasis on the influence of stress level
on shear strength and its significance for the stability of rockfill
slopes. Much of the chapter deals with laboratory and analytical work
carried out at the Building Research Establishment (BRE).
Rockfills are generally characterised by high shear strength,
particularly when heavily compacted into a dense state. Although shear
strength can be of significance in relation to earth pressure and load
carrying properties, it is the relationship between shear strength and
slope stability that is of major importance for rockfill embankments.
The shear strength of uncompacted fills can be of interest, but the
primary concern in this paper is with the strength of engineered
heavily compacted rockfills typically used in embankments for dams.
In-situ determination of strength is difficult and the emphasis is
on the laboratory measurement of the strength of recompacted samples.
The parameter of major interest is ~', the drained angle of shearing
resistance (sometimes referred to as the angle of internal friction).
It has been customary to express the result of a single drained test as
a value of ~' such that
(1)

where (cr'l/cr'3)max is the maximum principal effective stress ratio. It


is assumed that there is no cohesion in rockfills (c'=O). The drained
shearing resistance is strongly dependent on dilatancy and hence is a
function of relative density ID (sometimes termed density index),
stress level, stress history and strain conditions in the test.
Attempts have been made to correlate the shear strength of rockfills
with index and strength properties of the parent rocks (Barton and
Kjaernsli, 1981; Delgado Rodrigues et al, 1982; Veiga Pinto et al,
1987). Only limited success has been achieved, probably because of
the dominating influence of relative density on shear strength.
53
E. Maranha das Neves (ed.), Advances in Rock/ill Structures, 53-72.
1990 British Crown. Reproduced by permission of the Controller, IIMSo.
54

2. LABORATORY TESTS

Triaxial testing equipment is commonly used to measure shear strength


in drained compression tests. Strain rate should not greatly affect
the measured strength. End restraint can increase the apparent
strength of cohesionless samples. The effect decreases with increasing
height to diameter ratio and is of little significance with the usual
height to diameter ratio of 2:1.
Drained tests are often simplest to perform and most closely
resemble field conditions, but in some situations undrained testing may
be required. Two different failure criteria may be considered.
(a) (cr' 1 -cr' 3) reaches a maximum value
(b) (cr' 1Icr' 3) reaches a maximum value
The two conditions are reached simultaneously in a drained test, but
not in an undrained test.
The direct shear test is appropriate for testing free draining
rockfills and plane strain testing equipment also has been used to
measure drained shear strength parameters of rockfills.

3. STRAIN CONDITIONS

Strength testing is usually carried out under axisymmetric conditions


in triaxial testing equipment. In many field situations associated
with embankments, plain strain conditions apply. Consequently there is
considerable interest in relating the triaxial angle of shearing
resistance (~'ax) to the plain strain angle of shearing resistance
(!)l'ps).

Rockfills have been tested in large scale plane strain equipment


(Marsal et aI, 1967; Marsal, 1973). When the plane strain tests were
compared with triaxial tests at the same minor principal stress, it was
found that !)l' at failure was greater in the plane strain tests,
typically by as much as 8. AI-Hussaini (1983) carried out tests on
dense crushed basalt and found that a sample sheared under plane strain
conditions exhibited significantly higher shear strength and smaller
axial strain at failure than a comparable sample tested under triaxial
compression. The difference between the two types of test was less
significant with increasing confining pressure and decreasing density.
The use of triaxial test parameters for plane strain field situations
is therefore conservative and may offset any small overestimate of
strength due to scale or density effects.
Leonards and Frost (1988) attributed to Schmertmann an approximate
expression
!)l'ax = !)l'ps - [(!)l'ps-32)/3] (2)

for q,'ps > 32. It was suggested that if !)l'ax< 32, assume !)l'ax= q,'ps.
Rearranging equation (2)
!)l' ps = 1. 5 (q,' ax) - 16 (3)
55

In the Danish Code of Practice (Steensen-Bach, 1989)

$'ps = 1.1 $'ax (4)

Rowe (1969) examined the relationship between the shear strength of


sands in triaxial compression, plane strain and direct shear and
derived the relationship,

tan $'ds = tan $'ps cos $'cv (5)

where $'ds direct shear angle of shearing resistance,


$'ps plane strain angle of shearing resistance,
$'cv critical state angle of shearing resistance.

4. INITIAL POROSITY

At low confining pressures the initial density or porosity of the


triaxial sample has a major effect on the shearing resistance of the
rockfill. This is because dilatancy, which has a major influence on
the shear strength of coarse grained soils, is largely controlled by
stress level and the state of packing of the soil particles (described
by 10). The large scale triaxial tests of Marachi et al (1972)
indicated that at a confining pressure of 200 kPa an increase of 1% in
initial porosity reduced $' by 0.5.
Dense rockfills tested at low confining pressures exhibit a
maximum or peak strength that is associated with strongly dilatant
behaviour. This is followed by strain softening to a constant volume
or critical state strength $'cv. There is no evidence that this type
of brittle behaviour in rockfills leads to progressive failure unless
they contain shale or clay type materials. The effect is suppressed at
very large confining pressures.
For strength parameters measured in the laboratory to be
applicable to the field situation it is essential that tests are
carried out on samples compacted to an appropriate density.

5. CONFINING PRESSURE

When triaxial tests are carried out to derive shear strength parameters
for a stability analysis, it is important that appropriate values of
cell pressure are used. The stresses in the triaxial test should
correspond to the range of stress which will be encountered on
potential critical failure surfaces in the embankment slopes.
The shear strength of a rockfill is a function of the stress level
and stress history. In fill materials the stress history is usually
very simple and often the fill is normally consolidated. Typically the
Mohr failure envelope is a curve passing through the origin. In a
dense state the failure envelope shows marked curvature, while in a
loose state the failure envelope may be almost linear.
In general $' measured in drained tests on coarse grained soils
56

decreases with increasing confining pressure. Billam (1971) tested


dense samples of a limestone sand over a wide range of confining
pressures. With a'3=3 kPa, ~'=58.4 and with a'3=10 MFa, ~'=36.9.
An investigation has been carried out at BRE to measure the
influence of confining pressure on the angle of shearing resistance of
a number of well graded and heavily compacted rockfills at low and
medium confining pressures (Charles and Watts, 1980). Drained triaxial
compression tests have been carried out on 230 mm diameter samples with
a maximum particle size of 38 mm.
The relationship between shear strength and stress level can be
described by either a curved failure envelope or by some expression
relating ~' to a'3.

5.1. Curved failure envelope

The curved failure envelope determined from a series of tests at


different confining pressures can be approximated by a relationship of
the form

(6)

It is readily seen from equation (6) that


lf/a' = A(a,)b-l (7)

dlf/da' = Ab(a,)b-l (8)

Consequently the ratio (dlf/da')/(lf/a') = b at any normal stress.


Tests on rockfill have suggested that for heavily compacted samples
typically b = 0.75 (Charles and Watts, 1980). It should be noted that
whereas the parameter b is inde~endent of the units used for stress the
parameter A has dimensions [aJ ( -b).
Figure 1 presents the results of 230 mm diameter triaxial
compression tests carried out at BRE on recompacted samples of 38 mm
maximum size well graded sandy gravel and soft rockfill. The soft
rockfill was a mixture of sandstone, siltstone and mudstone. Although
this fill was formed from weak rock, the dense fill had high strength
at low confining pressures due to the high rate of dilation at failure.
When the results presented in Figure 1 are expressed in terms of
equation (6), the following values of A and b are obtained.

TABLE 1. Shear strength parameters for


some fills where If=A(a,)b (stress in kPa)
ID A b

Sandy gravel 0.95 4.4 0.81


Soft rockfill 0.95 4.2 0.75
Soft rockfill 0.70 1.4 0.90
57

1500.---------------------------------,
Fill type ID
SGD Sandy gravel 0.95
SRD Soft rockfill 0.95
SRM Soft rockfill 0.70

1000

7f
(kPa)

500

o 500
o' (kPa)

Figure 1. Shear strength of some rockfills measured in drained


triaxial compression tests on 230 mm diameter samples (~f va').

5.2. Relationship between ~' and a'


The result of each laboratory shear strength test can be expressed as a
value of ~' on the assumption that the cohesion intercept (c') is zero.
The results of a series of tests at different confining pressures are
then plotted as ~' versus log 0'3.

~' (9)

Expressions of this type have been proposed for rock fill by Barton and
Kjaernsli (1981) and by Bolton (1986) for sands. The difference
between ~' at a particular stress and some basic value of~' (~'B) is
expressed in terms of a logarithmic function of the effective stress.
Figure 2, which shows ~' plotted against log 0'3 for sandy gravel and a
S8

soft rockfill, is based on the same experimental data used to produce


the failure envelopes shown in Figure 1. Applying equation (9) (with
$'B = $'cv) to the results shown in Figure 2, the following values are
found for the constants Cl and C2 and $'cv.

TABLE 2. Shear strength parameters for some


fills where $' = $'cv + Cl 109(C2/ cr '3)

ID Cl C2 ~'cv

Sandy gravel 0.95 11.5 900 kPa 47


Soft rockfill 0.95 18 500 kPa 36
Soft rockfill 0.70 7 500 kPa 36

60

50

cJ>'
(degrees)

D~~
40
Fill type ID
[:, Sandy gravel 0.95

0 Soft rockfill 0.95

0 Soft rockfill 0.70

30
10 100 1000
CJ3 (kPa)

Figure 2. Shear strength of some rockfills measured in drained


triaxial compression tests on 230 rom diameter samples ($' v log cr'3).
59

Maksimovic (1989) has suggested the following expression for triaxial


compression tests,

(10 )

where $'B is the basic angle of shearing resistance, ~$' is the maximum
contribution of dilatancy, PTX is the value of 0'3 at which $'= $'B+
~$' /2. When 0'3= 0, $'= $'B+~$' and when 0'3~~, $'= $'B.
Usually a programme of tests is not carried out over a
sufficiently wide range of stresses to examine the validity of equation
(10). However Billam (1971) tested several materials in drained
triaxial compression tests over a range of confining pressures from 3
kPa to 10 MPa and his results on crushed anthracite have been used by
Maksimovik to illustrate the applicability of the above relationship.
This is shown in Figure 3. While it is clear that Maksimovik's
relationship does give a good approximation to Billam's data, the
experimental points could also be represented by a linear relationship
on the semi-log plot for confining pressures between 10 and 1000 kPa
which covers virtually the whole range of stresses which are of
practical interest.

50 I I I I
_ </J'o = --
.-........
48.6 0

.~ eBiI"m (1911) .
45 - I

\.<UBHr(l.;~x)
0
'<t
</J'
~
(degrees)
II
40 - '0-
-
<l

35 -
P TX =64.3kPa .~ -

</J'B = 32.2 0
- - . - .

30 I I I I
1 10
a 3 (kPa)
Figure 3. Shear strength of crushed anthracite ($' v log 0'3).
60

6. DILATANCY

The different methods of representing the relationship between shear


strength and stress level recognise that the shear strength has two
components.

6.1. A basic angle of shearing resistance

This is commonly identified with the critical state, steady state or


constant volume angle of shearing resistance ($'cv) but occasionally
with an angle of friction ($p).
6.2 A component due to dilatancy

Rowe (1962) developed stress dilatancy theory linking the principal


effective stress ratio with the rate of dilation. Figure 4 illustrates
the effect of dilatancy on shear strength by plotting $' as a function
of the rate of dilatancy at failure (dEv/dEl)f using the laboratory
tests on which Figures 1 and 2 are based. The results for a soft
rockfill lie on one straight line although tests were carried out over
a range of confining pressures at different initial porosities.

7. PARTICLE SIZE
With many coarse grained soils, particularly rockfills, it will not be
possible to test at full size. Some procedure is then required for
scaling down the material for testing. Sometimes a sample is tested
that has a grading parallel to that in the field. However this can
mean that the sample contains an excessive quantity of fine material.
The simplest approach is to remove the oversize material and test the
remainder. Clearly laboratory tests should be carried out at the
largest size possible to minimise the extent to which a sample has to
have its particle size distribution scaled down. In relating laboratory
results to field behaviour, the effect of particle size on the shear
strength should be considered.
The maximum particle size used in the BRE triaxial tests was 38 mm
whereas rock fill embankments may contain rock fragments with a
dimension of a metre or more. Charles and Watts (1980) reviewed the
data concerning the relationship between shear strength and maximum
particle size and concluded that it may often be a relatively minor
effect, providing that the scaled material is still behaving as a free
draining granular fill.
It should first be established that the ratio of the significant
dimension of the testing equipment (L) to the maximum particle size of
the rockfill (Dmax) is adequate. In triaxial testing it is usually
considered that L/Dmax should be a least 6 where the significant
dimension L is the diameter of the sample. A larger ratio might be
required when testing uniformly graded fills.
The ratio is likely to be much more critical in shear box tests.
In these tests the position of the failure plane is defined by the
61

70 r--------,--------,--------~--------_,--------._------_,

, /
60
/

(degrees)

50

40

D
r /'..

0
Fill type

Sandy gravel

Soft rockfill
ID
0.95
0.95

of D Soft rockfill 0.70

30
+0.5 0 -0.5 -1 1.5 -2 -2.5
( d Ev )
dE, I

Figure 4. Relationship between shear strength and dilatancy measured in


drained triaxial compression tests on 230 mm diameter samples.
62

apparatus whereas in triaxial tests the position of the failure plane


is much less well defined. A minimum ratio of L/Dmax of 10 should be
adopted where L is the height of the sample. Again a larger ratio may
be required for uniformly graded fills.
There is some evidence that ~'cv tends to increase with increasing
particle size. Direct shear tests in a 300 mm shear box on single size
crushed granite placed loosely showed ~' increasing from 34 for 0.15
mm particles to 39 for 4.5 mm particles (Lewis, 1956). Ohne et al
(1987) found that the angle of repose for a gravel increased from 46
to 57 as the mean particle size increased from 7 mm to 55 mm.
In so far as particle size effects are related to particle
breakage, it should be noted that contact forces in dense well graded
materials are small and particle breakage will be very low. Also well
rounded particles of gravel will be less prone to fracture than angular
fragments of rockfill. The results of direct shear tests on basaltic
rockfill caused Nieble et al (1974) to conclude that with uniform grain
sizes the shear strength decreased as the particle size increased, but
with well graded samples the effect was practically negligible.
When results are interpreted according to equation (9), the major
effect of particle size is probably associated with its relationship to
particle strength and hence C2. Particle size may have a smaller
effect on Cl which is the logarithmic gradient of reduction in shear
strength with increasing stress level.
The effect of particle size on shear strength is complex.
Generally it has been concluded that as particle size increases the
shear strength of compacted samples decreases. Marachi et al (1972)
came to this conclusion after testing in triaxial compression several
materials with parallel gradings and different maximum particle sizes.
However laboratory information comes from comparisons of tests on
samples with maximum particle sizes between 200 mm and 5 mm, whereas
the prediction of field performance may require knowledge of the
behaviour of particles with a maximum size of 1000 mm or more.
More information of the behaviour of rockfill in the field is
required. Gallacher (1972, 1988) described field tests on a gravel
fill at Megget Dam, Scotland, in a large 5 m x 2 m x 1.5 m deep tipping
box holding 34 t. The angle of repose for loose fill was 37.5, the
maximum slope for dense fill protected by a thin membrane was 54.
Blee and Riegel (1951) described large shear box test carried out on
sandstone fill at South Holston Dam which gave ~'=45.

8. SLOPE STABILITY
"Curiously, when dumped rockfill was used, steeper slopes were
automatic, and when compacted lifts began, accompanied by laboratory
based theorisation, there has been a flattening of slopes" (De Mello,
1977) .
Rockfill dams were built in California in the second half of the
nineteenth century with slopes as steep as 1 vertical in 0.5 horizontal
(Galloway, 1939). Embankments up to 30 metres in height were built of
rock fill dumped loosely in position. The construction of slopes
63

steeper than the angle of repose of the loose rockfill (which is


typically about 1 in 1.3) was achieved by hand placement of stone to
form a rubble retaining wall. In recent years different construction
techniques have been adopted for the construction of rockfill
embankments and during the last 30 years many major dams have been
built of heavily compacted rockfill. The use of modern heavy earth-
moving machinery has led to rockfill being placed in comparatively thin
layers (0.5 to 2 m deep) and compacted with heavy vibrating rollers.
This method results in very dense well graded rockfills with strength
and deformation properties greatly superior to those of the
uncompacted, uniform sized rockfills previously used. Yet, as de Mello
(1977) pointed out, slopes have tended to become flatter.
It would seem that some modern rockfill structures may have slopes
with high factors of safety that do not fully utilise the shear
strength properties of the fill. A quite small value for the factor of
safety obtained in a stability analysis (say, F = 1.4) may have been
based on conservative assumptions about the shear strength of the
compacted rockfill (say, angle of shearing resistance, $'=40).
The pronounced curvature, particularly at low stresses, exhibited
by Mohr failure envelopes for compacted rockfills has formed an
obstacle to the use of realistic shear strength parameters in stability
calculations as analyses have been based on linear Mohr failure
envelopes. Recently stability analyses have been carried out for
materials with curved failure envelopes (Maksimovik, 1979; Costa Filho
et al, 1982). It is now feasible both to measure the shear strength
parameters of a compacted rockfill and to use these realistic
parameters in slope stability analyses. A computer programme which
carries out the semi-rigorous Bishop (1955) circular arc stability
analyses has been modified to accept non-linear failure envelopes. A
dimensionless stability number has been used in conjuction with the
non-linear failure envelopes (Charles, 1982; Charles and Soares, 1984a
and 1984b) to facilitate the production of charts for the rapid
assessment of rockfill slope stability. These stability analyses give
an indication of the slopes required to achieve adequate factors of
safety in rockfill embankments of different heights.

9. STABILITY CHARTS FOR ROCKFILL SLOPES

The high perme~bility of rockfill ensures that in many rockfill slopes


there are no excess pore pressures. A dimensionless stability number
(r) can be calculated such that,
r = .u1!!l.(l-b) (11)
A

r is the same for geometrically similar slip surfaces in geometrically


similar slopes of the same uniform rockfill. Using the computer
programme, stability numbers have been calculated for the critical slip
surfaces of uniform rockfill slopes in which there are no pore water
pressures and which are built on strong rock foundations. Although the
64

shear strength parameter b is likely to be of the order of 0.75 for


compacted rockfill, a range of b values (0.5 < b < 1) has been
analysed. In Figure 5 the stability numbers r calculated from Bishop
stability analyses has been plotted against cot ~ for 0.5 < cot ~ < 2.0
and 0.5 < b < 1.0. For each value of b the relationship between rand
cot ~ is almost linear. When b=l, 1f=Ao' (this corresponds to a
constant angle of shearing resistance) and r=cot~. The positions of
the centres of the critical failure surfaces have been plotted in
Figure 6. As the value of b decreases from 1.0 to 0.5, the radius of
the critical slip surface also decreases ie the critical surface goes
deeper.

10. STABILITY OF SUBMERGED ROCKFILL SLOPES

When a rockfill slope is submerged by a rising reservoir level, the


factor of safety against stability failure will change. If the slope
is protected by an impermeable membrane water will not be able to
penetrate into the rock fill and in this situation submergence of the
upstream slope may increase the factor of safety against stability
failure. In the more important case where there is no impermeable
membrane on the slope, the water will penetrate and submerge the
rockfill itself. The effect of submergence on the stability of the
slope is more complicated in this situation. If the stability number r
has been determined for a slope in a uniform rockfill from the
stability chart, this will make it possible to calculate the factor of
safety for both the 'dry' slope prior to submergence (Fdry) and the
totally submerged slope (Fsub). (The term 'dry' slope is used to
denote a slope that is not submerged; the rockfill within the slope may
have a quite high moisture content.) In the former case the bulk
density of the rock fill (y) must be used in equation (11) and in the
latter case the submerged density (y').

(Fsub/Fdry) (y /y' ) (l-b) (12)

In general y/y' will be slightly smaller than 2. If y/y' = 2 and


b= 0.75 then, Fsub/Fdry = 1.19. (This simple analysis assumes that the
shear strength parameters A and b are not affected by submergence.)
Although the factor of safety of the fully submerged slope will
thus be slightly greater than that of the dry slope, partial
submergence can reduce the factor of safety. A minimum factor of
safety is generally obtained when the submerged depth is about 30% of
the height of the rock fill slope. Figure 7 shows the effect of a
rising water level on 50 metre high rockfill slopes with cot ~ = 1 and
2 respectively and with typical shear strength and density parameters.
The reduction in factor of safety due to partial submergence is greater
with flatter slopes. With cot ~ = 2.0, the minimum factor of safety as
the water level rises is almost 10% smaller than the factor of safety
for the dry slope.
65

5~--------------~--------------~--------------~

r 3 b= 0.8

b = 1 .0

1L-----------------~--------------~-------------------
0.5 1.5 2
cot p

Figure 5. Stability numbers for circular arc analyses.


66

cot ~

x
H 0

x (X,Y)
\',- '-
-1
\
\
\

3
\f \
\

.--
Y
H
2

lL-______- L_ _ _ _ _ _ _ _L -_ _ _ _ _ _- L_ _ _ _ _ _ _ _L -_ _ _ _ _ _ ~

0.5 0.6 0.7 0.8 0.9 1.0


b

Figure 6. Location of centres of critical slip surfaces.


67

11. DESIGN OF ROCKFILL SLOPES

The following procedure is suggested for the design of a slope to be


built of compacted rockfill to a particular height. It is assumed that
the rockfill is to be placed on a strong rock foundation, that it is
sufficiently permeable to prevent excess pore pressures being set up
and that it will not be submerged.
11.1. Determination of the rockfill shear strength parameters

Parameters A and b can be obtained from large scale drained triaxial


compression tests on samples of the compacted rockfill (Charles and
Watts, 1980). The use of this type of test has been justified in an
earlier section. It is important that tests are carried out on samples
compacted to a dry density and at a moisture content similar to that
which will be obtained in the field. Modern heavy vibrating rollers
can achieve very high field densities. It is also important to carry
out the tests at appropriate confining pressures. Figure 8 shows
(cr'm/yH), where cr'm is the maximum normal stress on the critical
failure surface, plotted against cot~. From this graph a preliminary
estimate can be made of the range of normal stress which will be of
interest in a particular situation. For example with b = 0.75 and cot ~
= 1.5, (cr'm/YH) = 0.31 and if y = 22 kN/m 3 and H = 50 m, then cr'm = 341
kPa. Triaxial tests at confining pressures of 30, 100 and 300 kPa
respectively would be appropriate in this situation. When the Mohr
circles have been plotted for the tests, the failure envelope can be
drawn. Generally this failure envelope can be described with
sufficient accuracy by equation (6), t f = A(cr,)b. It is seen therefore
that,

log tf = b log cr' + log A (13)

Consequently if the failure envelope is replotted on log-log graph


paper, a straight line should be obtained with a gradient equal to b.
The parameter A can also be obtained from this plot.
11.2. Selection of an appropriate factor of safety

In the design of slopes for embankment dams, values of F as low as 1.4


are sometimes quoted. However these factors of safety have usually
been calculated using conservative values for the shear strength
parameters of rockfills. In the present design method more realistic
values of rockfill shear strength parameters are being adopted and
therefore higher values of F may be appropriate. The very high shear
strength of compacted rockfill at low stresses is very brittle with a
large reduction in strength rapidly occurring once the peak strength
has been realised. Furthermore the peak strength will not be achieved
simultaneously at all points on a potential failure surface. The value
of F selected in a particular case will involve engineering judgement
and should depend on the importance of the rock fill structure and the
confidence with which the shear strength parameters have been
68

0.5

cot p =2

0~--~--~72----~--~

Figure 7. Effect of submergence on slope stability (A=5, b=O.75,


,,(/y'=2).

1.5 2
cot {3

Figure 8. Maximum normal effective stress on critical slip surface.


69

determined.

11.3. Calculation of the magnitude of the stability number

When the shear strength parameters A and b have been determined and a
factor of safety has been selected, the stability number can be
calculated using equation (11), r = IiIBl(l-b)
A

11.4 Determination of the slope angle ~

With the value of r calculated, the slope cot ~ can be ascertained from
Figure 5 using the stability numbers from the Bishop analysis.

12. EXAMPLES OF USE OF STABILITY CHARTS


Assume that tests on samples of compacted rockfill have shown that the
relationship t f = 5.0 (0')75 can adequately represent the shear
strength behaviour of the rockfill in the stress range appropriate to
the embankments. The embankments are to be built on firm rock
foundations, the rock fill is free draining and will not be submerged.
A factor of safety of 1.8 is required in both cases. The bulk density
of the rockfill is 22 kN/m 3

12.1. Example; 100 m high embankment


r = IiIBl(l-b) 1.8(22 X 100)25 2.47
A 5.0
from Figure 5, cot ~ = 1.32.
12.2. Example; 10 m high embankment
r = IiIBl(l-b) 1.8(22 X 10)25 1.39
A 5.0

from Figure 5, cot ~ 0.52.


12.3. Discussion
The first example illustrates how even a very high rockfill embankment
can be safely built with slopes almost as steep as the angle of repose
of loose uncompacted rockfill in certain circumstances. A weak
foundation, the presence of a wet clay core in a rockfill dam or the
submergence of the upstream slope would necessitate the use of flatter
slopes. The second example illustrates how small embankments of
compacted rockfill might be built with slopes steeper than the angle
of repose of the loose rockfill. In this latter situation it would be
necessary to contain and protect the surface of the slope.
70

13. CONCLUDING REMARKS

13.1. The shear strength of heavily compacted rockfill is strongly


dependent on placement density and stress level.

13.2. A simple relationship of the form 1 = A(o,)b can in general


adequately describe the failure envelope and the parameters A and b can
be determined from a series of drained triaxial compression tests
carried out on the compacted rockfill at appropriate confining
pressures.

13.3. Rational analysis of the stability of slopes of compacted


rockfill requires the use of realistic shear strength parameters based
on the curved failure envelopes typically found for compacted rockfills
at low and medium stresses.

13.4. The use of a non-dimensional stability number


r = IJrBl(l-b)
A

has made it possible to produce charts for the rapid assessment of the
stability of rockfill slopes.

13.5. The stability number also indicates how the factor of safety is
related to the height of the slope; for geometrically similar slopes in
the same uniform rockfill the factor of safety
F l/H(l-b)

13.6. Total submergence of a slope will typically increase the factor


of safety by nearly 20%. Partial submergence can however reduce the
factor of safety with a minimum value occurring with the water level at
a height of about 30% of the total slope height. The reduction is
greater with flatter slopes and is about 10% for a slope with cot ~ =2.

ACKNOWLEDGEMENT

The work described in this paper formed part of the research programme
of BRE and the paper is published by permission of the Chief Executive
BRE. Crown copyright 1990.

REFERENCES

Al-Hussaini M (1983). Effect of particle size and strain


conditions on the strength of crushed basalt. Canadian
Geotechnical Journal, vol 20, no 4, pp 706-717.
Barton Nand Kjaernsli B (1981). Shear strength of rockfill.
71

ASCE Journal of Geotechnical Engineering Division, vol 107, no


GT7, pp 873-890.
Billam J (1971). Some aspects of the behaviour of granular
materials at high pressures. Proceedings of Roscoe Memorial
Symposium on Stress-Strain Behaviour of Soils, Cambridge
University, pp 69-80. Fou1is, Henley.
Bishop A W (1955). The use of the slip circle in the stability
analysis of slopes. Geotechnique, vol 5, no 1, pp 7-17.
Blee C E and Riegel R M (1951). Rockfil1 dams. Transactions of
4th International Congress on Large Dams, New Delhi, vol 1, pp
189-208.
Bolton M D (1986). The strength and dilatancy of sands.
Geotechnique, vol 36, no 1, pp 65-78.
Charles J A (1982). An appraisal of the influence of a curved
failure envelope on slope stability. Geotechnique, vol 32, no
4, pp 389-392.
Charles J A and Watts K S (1980). The influence of confining
pressure on the shear strength of compacted rockfil1.
Geotechnique, vol 30, no 4, pp 353-367.
Charles J A and Soares M M (1984a). Stability of compacted
rockfill slopes. Geotechnique, vol 34, no 1, pp 61-70.
Charles J A and Soares M M (1984b). The stability of slopes in
soils with nonlinear failure envelopes. Canadian Geotechnical
Journal, vol 21, no 3, pp 397-406.
Costa Filho L M , Froes A Sand Romanel C (1982). Analise de
estabilidade de taludes em solos com envoltoria de resistencia
nao-linear. 3rd Latin American Conference on Numerical Methods
in Engineering, Buenos Aires.
Delgado Rodrigues J, Veiga Pinto A and Maranha das Neves E (1982).
Rock index properties for prediction of rockfil1 behaviour.
Proceedings of 4th International Congress of International
Association of Engineering Geology, New Delhi, vol 6, pp 39-47.
De Mello V F B (1977). Reflections on design decisions of
practical significance to embankment dams; 17th Rankine lecture.
Geotechnique, vol 27, no 3, pp 281-354.
Gallacher D (1972). A study of plane strain tests on granular
material. MSc thesis, Heriot Watt University, Edinburgh.
Gallacher D (1988). Asphaltic central core at the Megget Dam.
Transactions of 16th International Congress on Large Dams, San
Francisco, vol 2, pp 707-731.
Galloway J D (1939). The design of rockfill dams. Transactions of
American Society of Civil Engineers, vol 104, pp 1-24.
Leonards G A and Frost J D (1988). Settlement of shallow
foundations on granular soils. ASCE Journal of Geotechnical
Engineering, vol 114, no 7, pp 791-809.
Lewis J G (1956). Shear strength of rockfill. Proceedings of 2nd
Australia New Zealand Conference on Soil Mechanics, pp 50-52.
Maksimovik M (1979). Limit equilibrium for non-linear failure
envelope and arbitrary slip surface. 3rd International
Conference on Numerical Methods in Geomechanics, Aachen, vol 2,
pp 769-777.
72

Maksimovik M (1989). Nonlinear failure envelope for coarse grained


soils. Proceedings of 12th International Conference on Soil
Mechanics and Foundation Engineering, Rio de Janeiro, vol 1, pp
731-734.
Marachi N D, Chan C K and Seed H B (1972). Evaluation of
properties of rockfill materials. ASCE Journal of Soil Mechanics
and Foundations Division, vol 98, SM1, pp 9S-114.
Marsal R J (1973). Mechanical properties of rockfill. Embankment
Dam Engineering; Casagrande Volume (eds R C Hirschfeld and S J
Poulos), pp 109-200. Wiley, New York.
Marsal R J, de Arellano L R and Nunez A (1967). Plane strain
testing of rockfill materials. Proceedings of 3rd Panamerican
Conference on Soil Mechanics and Foundation Engineering,
Caracas, vol 1, pp 249-271.
Nieble C M, Silveira J F and Midea NF (1974). Some experiences on
the determination of the shear strength of rockfill materials.
Proceedings of 2nd International Congress of International
Association of Engineering Geology, Sao Paulo, vol 1, theme IV,
34.1-12.
Ohne Y, Tatebe H, Narita K and Okumura T (1987). Discussions on
seismic stability of slopes for rockfill dams. Proceedings of
International Symposium on Earthquakes and Dams, Beijing, vol 1,
pp 407-417.
Rowe P W (1962). The stress-dilatancy relation for static
equilibrium of an assembly of particles in contact. Proceedings
of Royal Society, vol 269, pp SOO-S27.
Rowe P W (1969). The relationship between the shear strength of
sands in triaxial compression, plane strain and direct shear.
Geotechnique, vol 19, no 1, pp 7S-86.
Steensen-Bach J 0 (1989). Strength evaluation of a natural sand.
Proceedings of 12th International Conference on Soil Mechanics
and Foundation Engineering, Rio de Janeiro, vol 1, pp 757-762.
Veiga Pinto A, Delgado Rodrigues J and Maranha das Neves E (1987).
Some improvements in the characterisation of basalts and
limestones for rockfill structures. Report no 679, Laboratorio
Nacional de Engenharia Civil, Lisbon.
CHAPTER 5
LABORATORY COMPRESSION TESTS AND THE DEFORMATION OF
ROCKFILL STRUCTURES
1. A. CHARLES

1. INTRODUCTION

Deformations of rockfill structures are usually related to laboratory


compression behaviour of rockfill measured in large oedometers. This
chapter deals mainly with experience within the United Kingdom and, in
particular, with field and laboratory measurements associated with
embankment dams and opencast mining backfills made by the Building
Research Establishment (BRE) during the last 20 years.
Rockfill structures are built for a variety of purposes which
include the following applications;
(a) embankment dams to retain reservoirs
(b) road embankments
(c) foundations for buildings
Deformations may occur within rockfill structures due to several
different causes. The most complex situations occur with embankment
dams. Compression of rockfill is associated with crushing of points of
contact and rearrangement of rock fragments and particles. An increase
in applied stress will cause compression due to these effects. An
increase in moisture content may weaken the parent rock material and
lead to further crushing; the beneficial effect of sluicing rockfill
was explained by Terzaghi (1960) as a softening of the particles and a
reduction in rock strength by saturation. Vibration may cause
compression by the particles being rearranged into a denser packing.

1.1. Changes in applied stress

These could arise in the following situations;


(a) stresses will increase during embankment construction due to the
weight of overlying layers of rockfill
(b) stresses will increase due to the weight of reservoir water acting
on an upstream membrane of a dam during reservoir impounding
(c) stresses will decrease due to the removal of the weight of
reservoir water acting on the upstream membrane of a dam during
reservoir drawdown
(d) vertical effective stress in the upstream rockfill will decrease
73
E. Maranha das Neves (ed.), Advances in Rockfill Structures. 73-96.
1990 British Crown. Reproduced by permission of tile Control/er, HMSo.
74

due to submergence of the upstream fill of a central core dam during


reservoir impounding
(e) total lateral stress of the upstream rockfill acting on a central
core of a dam will increase during reservoir impounding
(f) vertical effective stress in the upstream rockfill of a central
core dam will increase due to reservoir drawdown
(g) total lateral stress of the upstream rockfill acting on a central
core of a dam will decrease during reservoir drawdown
(h) vertical stress will increase due to the weight of a structure
built on a rockfill embankment

1.2. Increase in moisture content

This could include the following situations;


(a) inundation of unsaturated rockfill in the upstream shoulder of a
central core dam on first filling of the reservoir
(b) saturation of the downstream fill of a dam from reservoir water
seeping through or around the core
(c) downward percolation of surface water through the rockfill
(d) downward percolation of water leaking from road drains in a highway
embankment
(e) rise of ground water table in rockfill used to backfill opencast
mining workings or quarries

1.3. Vibrations associated with dynamic loading

compaction of rockfill during placement is usually achieved by a


vibrating roller. The better this compaction, the less vulnerable will
be the rockfill to volume reduction on subsequent dynamic loading which
could be associated with the following situations;
(a) seismic events
(b) traffic loading on a road embankment
(c) machine vibrations on a rockfill foundation
Compression due to this cause is not considered further in this
chapter.

2. LABORATORY ONE DIMENSIONAL COMPRESSION TESTING

Rockfill compressibility is most commonly measured under conditions of


one dimensional compression in an oedometer and the deformation of a
rockfill structure due to a change in stress or moisture content can
often be related to the behaviour of the rockfill measured in
appropriate oedometer tests. Rockfills contain material with large
particles and consequently large scale testing equipment is required.
Considerable numbers of tests have been carried out on rockfill
materials in large oedometers (Kjaernsli and Sande, 1963; Sowers et aI,
1965; Fumagalli, 1969; Marsal, 1973). A correction for side friction
has been described by Penman and Charles (1976).
The stiffness of rockfill is described by the constrained modulus
(D) where 0 is the ratio of an increment of applied vertical stress
75

(dO v ) to the increment of vertical strain (dEv) it produces. (The


constrained modulus is sometimes denoted by M, or may be described as
the oedometric modulus, Eoed). The stress-strain properties are
usually non-linear and the value of the constrained modulus will depend
not only on initial density and moisture content of the rockfill but
also on stress level and stress history.
Large oedometer compression tests on rockfills have led to the
following conclusions about the constrained modulus (Kjaernsli and
Sande, 1963; Sowers et aI, 1965);
(a) D increases with increasing hardness of the rock
(b) D increases with increasing relative density (ID)
(c) D is larger for smooth surfaced materials
(d) D is larger for a broad grading
(e) the major mechanism is crushing of highly stressed points of
contact between particles which in turn results in some reorientation
of particles
(f) the rate of continuing settlement is similar to the secondary
compression of clays
The compressibility of some heavily compacted fills (125 mm
maximum particle size) measured in a one metre diameter oedometer at
BRE is shown in Figure 1. The fills were compacted in layers with an
electric vibrating hammer. All the fills received a comparable degree
of compaction. The slate and sandstone rockfills were more
compressible than the sandy gravel. The sandy gravel was tested at two
other initial densities and the major influence of initial density on
compression was clearly indicated (Figure 2) .
The one dimensional compression behaviour of most rockfills on
first loading is non-linear and can often be approximated to a
relationship of the form

Ov (1)

The constrained modulus is therefore stress dependent and at any


particular stress there is a secant modulus (Dsec) and a tangent
modulus (Dtan) such that when equation (1) is a reasonable
representation of stress-strain properties

Dsec (2)

Dtan (3)

Figure 3 shows aO. 5 plotted against relative density (ID),


sometimes termed density illdex, for a sandy gravel fill and a sandstone
rockfill. At a given relative density the constrained modulus of the
sandy gravel was typically 4 times as large as the constrained modulus
of the sandstone rockfill.
In some cases the unloading and reloading stress strain
relationships can be of interest. This will be the case for
deformations in the upstream rockfill of a central core darn due to
reservoir draw down and refilling. The constrained modulus is very much
greater in these situations. This is illustrated in Figure 4 which
76

av (kPa)
o 500 1000 1500
o

Sandy gravel
a

(%)

'Yd D'YH D'


kN/m 3 MPa MPa
Sandy
gravel 21.5 111 136
Slate 21.9 46
rockfill 38
6 Sandstone 26
rockfill 20.2 22

'-------------'------------~-- .. -~----------'

Figure 1. Compressibility of heavily compacted rockfills measured in


one metre diameter oedometer.
DYH = secant constrained modulus with crv=yH
D* = constant equivalent constrained modulus
(crv=yH at points marked "a" for embankment dams built of these fills)

shows the initial loading and unloading of three samples of sandstone


rockfill ("a", "b" and "c") compacted to different initial densities.
Also shown is the reloading ("ar") of the least well compacted sample
"an which is much stiffer on reloading than the initial loading of any
of the three compacted samples.
Poorly compacted unsaturated rockfill is likely to be susceptible
to collapse compression when first inundated (section 8). This is
associated with softening and reduction in strength of the rock
particles. The effect of inundation on an unsaturated rockfill can also
be examined in an oedometer test.
77

o 500 1000 1500

10

12

Figure 2. Compressibility of sandy gravel measured in one metre


diameter oedometer on samples compacted to different initial densities.
(a) Yd 21.5 kN/m 3 , Io 0.82
(b) yd 21.2 kN/m3 , Io 0.77
(c) yd 17.4 kN/m3 , Io 0.0
3000

1o

(kPa)O.5

2000

1000

o
Yo
O/sandstone
O~ rockfill
-0---
o 0.5
'0

Figure 3. Influence of relative density on compressibility of


sandstone rockfill and sandy gravel.
crv aE v 2
ID = relative density
79

av(kPa)

o 1000 2000 3000


~---------------,----------------,----------------,--------,

ar

fd
Test ID
(kN/m 3 )

a 16.4 0.24
d'2. ar 19.7 0.84
> 10
""' b 17.5 0.46
c 19.3 0.78

15

20~--------------------------------------------______~

Figure 4. Compressibility of poorly compacted sandstone rockfill


during loading, unloading and reloading.
80

3. FIELD PROPERTIES

Information about rockfills used in some UK embankment dams is


presented in Tables 1 and 2 below. Table 1 gives an indication of
previous practice in rockfill placement in terms of layer thickness,
the weight of the vibrating roller (VR) and the number of passes.
Table 2 gives a broad indication of average porosity (n) and air voids
(AV). It was not possible to include all the relevant details in the
two tables. The layer thickness and method of compaction may have been
changed during construction of the embankment, some water may have been
added to the rockfill, there may have been more than one type of
rockfill used in the embankment. Furthermore some of the quoted
density results may be less reliable than others, being based on far
less test data or incomplete information.

TABLE 1. Field placement of rockfill

Dam Date Rockfill Layer Pass VR Wt


m no t
Balderhead 1965 shale 0.8 2 8.5
Scammonden 1969 sandstone/mudstone 0.9 5 11.5
Llyn Brianne 1971 mudstone 0.9 4 13.5
Winscar 1974 sandstone 1.7 4 13.5
Marchlyn 1979 slate 1.0 4 13.5

In addition to information on densities achieved with heavily


compacted rockfills in embankment dams, some information has been
provided in Table 2 for two rockfills used as opencast mining
backfills. At Horsley the backfill consisted mainly of mudstone and
sandstone; at Tranent of mudstone, siltstone and sandstone. Neither
fill was systematically compacted. The densities were determined from
small samples and can only be a crude guide to the condition of these
two very variable fills.

TABLE 2. Field density of rockfill

Location Structure pd w ps n AV
Mg/m 3 % Mg/m3 % %

Balderhead embankment dam 1. 94 9 2.66 27 10


Scammonden embankment dam 2.02 7 2.69 25 11
Llyn Brianne embankment dam 2.35 3 2.75 15 7
Wins car embankment dam 2.03 6 2.60 22 10
Marchlyn embankment dam 2.25 4 2.81 20 11
Horsley opencast b' fill 1. 70 7 2.54 33 21
Tranent oEencast b'fill 1.54 9 2.45 37 23
[pd = dry density of fill, ps particle density (specific
gravity) , w = moisture content]

The five dams had porosities in the range 15% to 27% and this
81

could be regarded as typical for modern well graded heavily compacted


rockfills. In contrast the two opencast backfills had much higher
porosities, 33% and 37% respectively.

4. FIELD MONITORING OF DEFORMATIONS


Deformations have been monitored at a number of UK rockfill dams during
construction, reservoir impounding and in the long term by BRE. Some
basic information about the rockfill used in these dams has been
summarised in tables 1 and 2.
Scammonden Dam in west Yorkshire, completed in 1969, is 73 m high
with an unusually wide crest which carries a motorway. It has an
upstream sloping soft clay core. The rockfill was a Carboniferous
sandstone with three mudstone zones in the middle of the embankment.
The clay for the core was placed at a moisture content well above
Proctor optimum. BRE measurements have been described by Penman et al
(1971) and Charles (1973).
Llyn Brianne Dam is a 90 m high rockfill dam situated in central
Wales. Embankment construction was completed in 1971. It has a wide
central clay core placed wet of optimum moisture content. The rock fill
is a slaty mudstone of the Lower Palaeozoic period and water was added
during placement. BRE measurements have been described by Penman and
Charles (1973a and 1973b) .
Winscar Dam in west Yorkshire was completed in 1974 and is 53 m
high. It has an upstream membrane of asphaltic concrete. The rock at
Wins car is the same Carboniferous series found at Scammonden. Water
was added to the rock fill during placement. The upstream asphaltic
membrane was placed after embankment construction was completed. BRE
measurements have been described by Penman and Charles (1985a).
Marchlyn Dam in north Wales forms the upper reservoir of the
Dinorwig pumped storage scheme. It was formed by building a rockfill
embankment on a glacial moraine and then placing the upstream membrane
over both rock fill and moraine. It is approximately 72 m high from toe
to crest. BRE made no measurements during construction but developed a
special inclinometer to measure delections from the 1 in 2 slope of the
membrane during impounding. This instrumentation has been described by
Penman and Hussain (1984).
Megget Dam in southern Scotland is 56 m high and has a central
bituminous core supported by gravel shoulders. The well graded gravel
was placed in 0.4 m thick layers and compacted with 4 passes of a 5.5
tonne vibrating roller. BRE measurements have been described by Penman
and Charles (1985b).

5. CONSTRUCTION DEFORMATIONS OF EMBANKMENTS


The settlement of rock fill structures during construction is due to the
self weight of the rockfill. Constrained moduli for fills from several
UK dams have been measured in a one metre diameter oedometer and it has
been found that the maximum settlement during construction can be
82

related to the constrained modulus determined in a large oedometer


test. This method of deformation prediction depends on the assumption
that the compressibility of the rockfill used in the field can be
measured in a laboratory test on 125 mm down material compacted to the
field density. The evidence from the application of the method to a
number of dams is that the use of the fraction below 125 mm does not
introduce major errors and that the lack of lateral restraint at the
slopes of the embankment has only a minor effect on the settlement
within the embankment provided that there is a reasonable factor of
safety against slope instability.
Although the stress strain properties of rockfills determined in
one-dimensional compression tests are generally non-linear, it has been
shown that the internal distribution of settlement during the
construction of a thick layer possessing self-weight can be predicted
with little error by the use of a constant equivalent compressibility
(Penman et aI, 1971). The corresponding constant equivalent value of
constrained modulus is denoted by D* and corresponds to the value
required to give the correct final displacement at mid-height for an
embankment of infinite extent. The method of calculating D* is
described in Appendix A.
For rockfill at a particular location within an embankment, the
value of D* will be a function of the one dimensional compression curve
and the final height of the rockfill column at that particular
location. Thus in a finite element analysis of the construction of a
homogeneous rockfill embankment, each element of rockfill is assigned a
value of D* which depends on the final height of the rockfill column at
that particular location and which is not varied as the analysis models
the construction of the dam in layers. This method of predicting
construction deformations was developed and successfully applied to
Scammonden Dam (Penman et aI, 1971) and Llyn Brianne Dam (Penman and
Charles, 1973a).
The constructional settlements of rockfill under self weight
measured on the centre-line at the highest sections of four dams are
plotted in Figure 5. The settlements have been adjusted to remove
foundation settlement. Charles and Penman (1988) have shown that the
maximum constructional settlement (Smax) is related to D* by an
empirical relationship

(4)

where y is bulk density of rockfill and H is height of embankment.


This is a little larger than the 0.25 (yH2/D*) which would
theoretically be expected in a wide embankment. The difference could
be due to either the absence of lateral restraint at the slopes of the
embankments or the difference between the compressibility of the 125mm
maximum size rockfill tested in the oedometer and the compressibility
of the larger material in the field. Both these effects tend to
increase the settlement. Rockfill dams usually have narrow crests and
the deformations, even on the centre-line, will not be strictly one
dimensional and some lateral spreading of the embankment will occur as
embankment construction proceeds. It is suggested that the semi-
83

empirical relationship given in equation (4) may be useful during the


preliminary design stage, prior to making a detailed finite element
analysis.

~
w /'

Smax
/'
(m)
05 /'
/'
Ma
/'

/.Me

0/
0 2 3
'Y H2 (m)
D*

Figure 5. Relationship between observed maximum construction


settlement (Smax) and constrained modulus (D*) measured in one metre
diameter oedometer.
(Ma) Marchlyn Dam (S) Scammonden Dam
(Me) Megget Dam (W) Winscar Dam
6. MOVEMENT OF UPSTREAM MEMBRANE EMBANKMENT DAMS DUE TO RESERVOIR
IMPOUNDING

The reservoir water applies a loading normal to the membrane. First


filling of the reservoir leads to major stress increases and embankment
deformations. Subsequent fluctuations in reservoir level affect the
stresses in the rockfill but have a much smaller effect on embankment
deformations as the rockfill is much stiffer under these unloading and
reloading stress changes (Figure 4).
The constant equivalent compressibility method works well for the
prediction of constructional movements because the stress changes
during construction generally correspond to a simple increase in mean
84

effective stress while the principal effective stress ratio remains


reasonably constant (Charles, 1976). However during reservoir
impounding the stress changes are more complex. In the rockfill
immediately beneath the membrane or immediately downstream of a
bituminous core, the mean effective stress increases when thp TJater
load is transmitted to it, but the principal effective stress ratio
tends to reduce as the water load acts in a direction close to the
direction of the minor principal stress during construction. Field data
has shown that the modulus during first water loading is typically
about twice as large as that operating during construction. This can
be explained by considering the change in principal stress direction
during these two different phases of loading (see Appendix B) .
The deformations due to first filling of the reservoir have been
monitored at Winscar and Marchlyn dams. The measured deflections
normal to the upstream membrane of bituminous central core are shown in
Figure 6. At Winscar vector movements were measured using horizontal
plate gauges. At Marchlyn the deflections of the membrane were
monitored with an inclinometer. It should be noted that as the
movements at Winscar were much larger than at Marchlyn, different
movement scales have been used in Figure 6 for the two dams.
Finite element analyses were carried out to predict deformations
during impounding at Winscar with parameters based on the assumption
that the bulk modulus had the same magnitude during reservoir filling
as it had during embankment construction and Poisson'S ratio was zero
during reservoir filling (see Appendix B). This approach gave good
predictions of movements normal to the upstream membrane except close
to the crest (Penman and Charles 1985a) .
In Figure 7a the maximum normal deflection of the membrane on
first filling the reservoir (nmax) is plotted against Smax from
published information for upstream membrane dams in all parts of the
world. Results from concrete faced rockfill dams have been included as
well as results from dams with bituminous membranes. In some cases
these records may show considerable variations due to a number of
factors;
(a) whether measurements were made immediately at the end of
construction and reservoir filling or some months or even years later
(b) the rate of reservoir filling and any lowering before reaching top
water level
(c) large foundation settlements may have occurred and not been allowed
for in the quoted movements
(d) differences in behaviour caused by steep valley sides
(e) non-uniformity of fill within the embankment cross-section
(f) differing upstream slope angles
(g) inaccurate or too few measurements
From these field measurements Charles and Penman (1988) derived an
empirical relationship between the maximum deflection of the membrane
on first filling the reservoir and the maximum settlement of the
embankment during construction

nmax = 0.25 Smax (5)


85

o. 200
,"
400mm

Figure 6. Membrane deflections during first filling of reservoir.


(W) Wins car Dam
(Ma) Marchlyn Dam
(Me) Megget Dam
86

Obm ocf 6 be

Me
02 0
'r'J/ 02

/
G
A
/
0

F
n max
(m)
Bi0 /
h max
(m)
6

e V
/
s7
0

01 0 01
/
/LP
l~~
D
0 / FA
B'bo/' 04
/
/
/6

/ Ma
0 2 4
~e
6

/
0

0
/ 0
0 05 0 05
smax (m) smax (m)

Figure 7. Relationship between membrane deflection during first


filling of the reservoir and construction settlement.
(a) nmax versus Smax
(b) hmax versus Smax
(Smax) maximum settlement measured during embankment construction
(nmax) maximum normal deflection of upstream membrane measured
during first filling of reservoir
(hmax ) maximum horizontal movement of central core measured during
first filling of reservoir
(bm) bituminous upstream membrane
(bc) bituminous central core
(cf) concrete faced
(AA) Alto Anchicaya (G) Golillas
(Ba) Bastyan (LP) Little Para
(Bi) Bigge (Mc) Mackintosh
(e) eethana (Me) Megget
(D) Dhunn (Mu) Murchison
(F) Finstertal (S) Salvajina
(FA) Foz do Areia (V) Venemo
(Ma) Marchlyn (W) Wins car
87

7. MOVEMENT OF CENTRAL CORE EMBANKMENT DAMS DUE TO RESERVOIR IMPOUNDING

Where the embankment has a central flexible diaphragm of asphaltic


concrete, the situation during first filling of the reservoir is
complex. The reservoir will submerge the upstream fill and, if it has
not been heavily compacted, some collapse compression may occur on
first filling of the reservoir. Submergence of the upstream fill will
decrease the vertical effective stress in the fill, but the total
horizontal pressure acting on the membrane will increase. At Megget
Dam it was found that the increase in horizontal thrust due to
reservoir impounding was 0.29 ywhw, where yw is the density of water
and hw was the reservoir head at the location where the horizontal
stress was measured.
In Figure 7b the horizontal deflection of the core (h max ) on first
filling of the reservoir is plotted against Smax for some bituminous
core dams. Megget and Dhunn have vertical cores while Finstertal has a
core that slopes upstream at 0.4 horizontal to I vertical. The amount
of information is rather sparse for any firm conclusions to be made
about a correlation between hmax and Smax. However from the available
data a relationship hmax = 0.4 Smax is tentatively proposed.

8. MOVEMENT OF ROCKFILL STRUCTURES DUE TO COLLAPSE COMPRESSION

Inundation of poorly compacted unsaturated rockfill will generally


produce some collapse compression. If an embankment dam has an
internal core, the reservoir will submerge the upstream fill and may
cause some collapse compression on first filling of the reservoir. In
a modern heavily compacted rockfill placed at a suitable moisture
content collapse compression is likely to be negligible. Settlement of
the upstream and downstream rockfill shoulders during first filling of
the reservoir at Llyn Brianne Dam have been compared; the difference in
settlement was very small and it was concluded that collapse
compression was negligible (Charles, 1987).
Building developments and road construction in the UK are
increasingly taking place on opencast mining backfills. Many opencast
coal mining sites now go well below the water table and problems due to
ground water during mining have increased. Field investigations have
demonstrated that ground water is a major cause of settlement of
opencast backfills. The settlement can occur at depth as the
groundwater table rises or close to ground level as surface water
penetrates into the fill through shallow excavations through the
surface crust. The implications of these results for building on
opencast backfill may be serious. If opencast backfill was placed
without systematic compaction and has not yet been inundated, it is
probable that it will be susceptible to collapse compression on
inundation. In this case it is not necessarily sufficient to allow a
certain period of time to elapse between restoration and development.
The effect of a rising ground water table on the settlement of an
opencast coal mining backfill has been investigated at Horsley in
Northumberland (Charles et al 1977, 1984; Charles and Burford, 1987).
88

Backfilling of different parts of the site took place between 1961 and
1970 with restoration completed in 1973. The fill was composed of
mudstone and sandstone and had been placed using truck and shovel and
dragline without any systematic compaction. Some information on
average fill density is given in Table 2 but wide variations occurred.
Pumping had been continued after the completion of backfilling and
had kept the water table below the level of the fill. Pumping stopped
in 1974 and the water table rose 34 m reaching a new equilibrium level
in 1977. In the period when pumping continued, settlements were very
small. When pumping stopped large movements occurred, the largest
settlement recorded so far (between 1973 and 1986) is 0.75 m. As the
ground water table rose, vertical compressions locally were as large as
2% but the average settlement measured over the full depth of inundated
backfill was smaller than 1%. Figure 8 shows the settlement measured
at the deepest part of the site. Although most of the measured
settlement can be attributed to the rising water table, significant
compression has also occurred well above the equilibrium water level.
Part of the site had been temporarily pre-loaded with a 30 m high
surcharge of fill and this greatly reduced subsequent settlement due to
the rising ground water table at this location.

74 75 76 77 78 79 80 81 82 83 84 85 86
O~~-,----,---,----.---.----'---'----.---.----.---.----r---'

100
-O'\.\~I:]...'V'V_'V_v_v_'V_v_v_v__ v_v_,, __,,__ v_7_v---v--v---"
E o~;i'..:~__"__"__ "_" __"____ , _ ---,,----,,---,,----,,------,,---~ --,,----,,----,,------"
E 200 \OO:D~o.
c:
(I) '0 \.o,,6~<>_<>-o--o-o-o--o--o---o---o-o--o
o o~o, 9

.,t300 \ '0--0",'0_____ 10
Q; "0 0 - - - - 0 - - - 0 - ___ 0 _ _ _ _ 0 _____ 0 - - - - 0 - - _ 0 _ _ _ _ _ 0
Vl
"'0-0
400
~o
--""""0
500 -0-0_0__0_ 1 _1- 0 - - 0 _ 0 _ _ _ _ _ 0

O,,--.---.--,,--.---,--.-~:~~.---.--.---.--.---,--,
11
10
_0-0-0-0 .2... 0-0 90--0--0--0---0-0---0
........ 0- 0 -:-
B-; ~~o 8

-
ON~

d'~
:0 Backfill
0.-
(I)
'" 40 /
/0
Cl
00
60 oJ

Natural
ground

Figure 8. Effect of rising ground water level on settlement of


opencast mining backfill at Horsley, 1973-1986.
89

9. CREEP SETTLEMENT OF ROCKFILL STRUCTURES

When rockfill construction has been completed and water level


fluctuations have ceased, the important changes of stress in the
embankment and its foundation have occurred. Further movements may be
associated with creep at constant effective stress. A relationship of
the following form can be used to predict the rate of creep settlement

s = a H 109(t2/tl) (6)

where s is the settlement of the crest of a rockfill structure of


height H which has occurred between times t2 and tl from the end of
construction. Values of a measured in laboratory oedometer tests are
generally significantly smaller than those observed in the field.
Sowers et al (1965) reported the long term settlement of many
rockfi11 dams in the United States. When settlement was plotted
against the logarithm of time since the middle of the construction
period, approximately linear relations were obtained. Values of a
ranged from 0.2% to 1.0%. The amount of settlement expressed as a
percentage of dam height did not appear to be related to rock type,
form of construction (eg upstream membrane, central core) or height.
The significant factor was the method of placement of the rockfill. The
dam with the greatest settlement had been dumped with limited sluicing,
the dam with least settlement had been compacted by rolling while
sluiced.
The settlement of the crest of the 73 m high Scammonden Dam
embankment composed of heavily compacted sandstone and mudstone
rockfill, is shown plotted against log time in Figure 9 (Penman, 1971;
Charles, 1973). There is an approximately linear relationship between
s and log t corresponding to a = 0.17%. Also shown in Figure 9 are the
settlements of two points within the embankment on its centre-line at
0.25 and 0.64 of the full height of the embankment.
Settlement has been monitored within the downstream shoulders of
several UK rockfill dams and creep deformations have been examined at
different depths in regions of the embankments not seriously affected
by reservoir fluctuations. Compression has increased linearly with the
logarithm of time since the end of construction as predicted. The
logarithmic creep rate parameter a is stress dependent for these
heavily compacted fills. Some results for Scammonden Dam and Llyn
Brianne Dam are given in Figure 10. Information about these dams is
presented in Tables 1 and 2 and in section 4.
At Scammonden Dam, with 0 < Ov < 1.3 MFa (figure lOa)

a = 0.13 ov (7)

where a is in % and Ov is in MFa.


At Llyn Brianne Dam, with 0 < Ov < 1.3 MFa (figure lOb)
a = 0.12 Ov (8)

where a is in % and Ov is in MFa.


90

o~~---------------------------------------------------

100
E
E
....C
Ql
E
Ql
-.::;
....Ql
Ul

200

L -_ _ _ _-L-_ _

300~ ____~__~__~L-~-LLL_ _ _ _ _ _- L -_ _~~~~-L~~_ _ _ _~


0.1 10
Time since end of construction (years)

Figure 9. Long term settlement of 73 m high heavily compacted


sandstone and mudstone rockfill embankment.

At Megget Dam (a sandy gravel fill), with 0 < Ov < 1.0 MPa
(l = 0.04 ov (9)

where (l is in % and ov is in MFa.

The settlement of opencast backfills has provided information


about the settlement of rockfills that have not been systematically
compacted. To reduce settlement of a 1.4 km section of trunk road, the
Tranent by-pass, which was to be built across the Blindwells opencast
coal mine near Edinburgh, the top 16 m of the 60 m deep mudstone,
siltstone and sandstone backfill was systmatically compacted. BRE has
measured settlement since 1984, both where the upper backfill has been
compacted and in an area of uncompacted backfill (Charles and Burford,
1987). During the first five years of monitoring, settlements of up to
91

0.3 ~------------------------------------------------,
(a) 73m high embankment of heavily
compacted sandstone and mudstone rockfill
o

o
0.2

ex
(%)

0.1

o 0.5 1.0 1.5


Uv (MPa)

U v (MPa)

Figure 10. Relationship between logarithmic creep paparneter a and


stress level (ov) for heavily compacted rockfill.
92

0.15 m have been measured on the compacted backfill and 0.28 m on the
uncompacted backfill. a for the uncompacted fill is about 1%.

10. CONCLUDING REMARKS


10.1. Deformations may be caused by changes in stress, changes in
moisture content and by dynamic effects.

10.2. Large oedometer tests can be used to provide suitable deformation


parameters to describe rockfill settlement behaviour during embankment
construction. A simple empirical relationship has been proposed
relating the maximum settlement during embankment construction (Smax)
to the constrained modulus measured in a large oedometer test (D*)

10.3. The maximum normal deflection of an upstream membrane (nmax)


during first filling of the reservoir can be related to the maximum
constructional settlement (Smax)

nmax = 0.25 Smax


10.4. Collapse compression is likely when a poorly compacted
unsaturated rock fill is inundated. Heavy compaction during placement
in layers largely eliminates this hazard with rockfill placement
moisture contents typically used in the UK.
10.5 Long term movements under conditions of constant stress and
moisture content can be related to a logarithmic creep rate parameter
a. For poorly compacted rockfills, a may be of the order of 1%. For
heavily compacted rockfills, a is likely to be much smaller and stress
dependent.

ACKNOWLEDGEMENTS
The work described in this paper was carried out as part of the
research programme of BRE and this paper is published by permission of
the Chief Executive BRE. Field measurements have been made with the
permission of, and in cooperation with, dam owners and their consulting
engineers: Scammonden and Wins car Dams - Yorkshire Water, Llyn Brianne
Dam - Welsh Water, Marchlyn Dam - Central Electricity Generating Board,
Megget Dam - Lothian Regional Council. Measurements at Horsley and
Blindwells opencast mining sites have been made in cooperation with
British Coal Opencast Executive.
93

12. APPENDIX A
CONSTANT EQUIVALENT CONSTRAINED MODULUS

The one-dimensional compression properties of most rockfills are


non-linear. It is useful to be able to calculate a constant equivalent
constrained modulus (D*) which will predict the construction settlement
at mid-height of an embankment of infinite extent. It has been shown
that the value of D* which gives the correct value of settlement at
mid-height will also predict construction settlement at other heights
within the embankment to a reasonable accuracy.
For an embankment of height H and constructed of rockfill of bulk
weight density y, D* can be determined as follows
D* (10)

where E* is an average strain due to an increase in stress of 0.5yH. E*


can be calculated to an acceptable degree of accuracy from the measured
laboratory one-dimensional compression behaviour using the following
expression

E* = 0.2 :t{Ecr-O.95YH + Ecr-O.85yH + Ecr-O.75YH + Ecr-O.65YH + Ecr=O.55yH -


Ecr=O.45YH - Ecr-O.35yH - Ecr-O.25yH - Ecr-O.15YH - Ecr-O.05YH} (11)

where Ecr-O.95YH is the vertical strain measured in the oedometer test


under a vertical stress of 0.95yH etc.
Theoretically the maximum construction settlement (Smax) of a wide
embankment of height H built of fill with a constant equivalent
constrained modulus D* will occur at mid height and is given by
smax = 0.25 (yH2/D*) (12)
In practice it has been found that typically Smax 0.30 (yH 2 /D*) for
large rockfill embankment dams.
If the stress-strain relationship measured in the oedometer is of the
form crv aE v2
D* 1.28 DyH = 1.28 (ayH)O.5 (13)

where DyH is the secant constrained modulus with crv yH.


Generally it is found that 1.0 < D*/DYH < 1.28.

13. APPENDIX B
STRESS PATHS DURING CONSTRUCTION AND RESERVOIR IMPOUNDING FOR UPSTREAM
MEMBRANE DAMS

It has baen found that the modulus of deformation controlling movements


94

observed on first filling the reservoir is significantly greater than


that governing construction deformations. To investigate this situation
small scale triaxial tests were carried out on a sandstone rockfill
(Charles, 1973, 1976). The tests followed a simplified version of the
field stress paths. In stage 1 of the tests the principal effective
stress ratio was kept constant while the mean effective stress was
increased. In stage 2 the major principal effective stress was kept
constant while the minor was increased. The main findings were as
follows.
(a) The relationship between mean effective stress and volumetric
strain was similar in the two stages of the test.
(b) The relationship between shear stress and shear strain was quite
different in the two stages of the test.
In explaining these findings it should be noted that the mean
effective stress was increasing in both stages while the shear stress
increased in the first stage and reduced in the second. Thus a similar
bulk modulus might be expected in the two stages, but the shear modulus
would be very different because in terms of shear stress the first
stage was a loading situation but the second stage was unloading. It
should also be noted that in the second stage of the tests the strain
increment normal to the applied stress increment was close to zero and
thus corresponded to a situation in which Poisson's Ratio was zero. If
it is assumed that the bulk modulus is the same in the two stages and
that Poisson's Ratio is 0.33 in stage 1 and zero in stage 2, it can be
shown that the Young's Modulus in stage 2 will be 3 times as large as
the Young's Modulus in stage 1 and the constrained modulus in stage 2
will be 2 times as large as the constrained modulus in stage 1.

REFERENCES
Charles J A (1973). Correlation between laboratory behaviour of
rockfill and field performance with particular reference to
Scammonden Dam. PhD thesis, University of London.
Charles J A (1976). The use of one-dimensional compression tests
and elastic theory in predicting deformations of rock fill
embankments. Canadian Geotechnical Journal, vol 13, no 3, pp
189-200.
Charles J A (1987). Discussion contribution on Collapse
compression of fills. Proceedings of 9th European Conference on
Soil Mechanics and Foundation Engineering, Dublin, vol 3, pp
1079-1080.
Charles J A, Naismith W A and Burford D (1977). Settlement of
backfill at Horsley restored opencast coal mining site.
Proceedings of Conference on Large Ground Movements and
Structures, Cardiff, pp 229-251. Pentech Press, London.
Charles J A, Hughes D B and Burford D (1984). The effect of a
rise of water table on the settlement of backfill at Horsley
restored opencast coal mining site, 1973-1983. Proceedings of
3rd International Conference on Ground Movements and Structures,
Cardiff, pp 423-442. Pentech Press, London.
95

Charles J A and Burford D (1987). Settlement and groundwater in


opencast mining backfills. Proceedings of 9th European
Conference on Soil Mechanics and Foundation Engineering,
Dublin, vol 1, pp 289-292.
Charles J A and Penman ADM (1988). The behaviour of embankment
dams with bituminous watertight elements. Transactions of 16th
International Congress on Large Dams, San Francisco, vol 2, pp
693-705.
Fumagalli E (1969). Tests on cohesionless materials for rockfill
dams. ASCE Journal of Soil Mechanics and Foundations Division,
vol 95, January, SM1, pp 313-330.
Kjaernsli B and Sande A (1963). Compressibility of some coarse
grained materials. Proceedings of European Conference on Soil
Mechanics and Foundation Engineering, Wiesbaden, pp 245-251.
Marsal R J (1973). Mechanical properties of rockfill. Embankment
Dam Engineering - Casagrande Volume (ed Hirschfeld R C and
Poulos S J), pp 109-200. Wiley, New York.
Penman ADM (1971). Rockfill. Building Research Station Current
Paper 15/71. BRE, Garston, Watford.
Penman ADM, Burland J B and Charles J A (1971). Observed and
predicted deformations in a large embankment dam during
construction. Proceedings of Institution of Civil Engineers,
vol 49, May, pp 1-21.
Penman ADM and Charles J A (1973a). Constructional deformations
in rockfill dam. ASCE Journal of Soil Mechanics and Foundations
Division, vol 99, SM2, pp 139-163.
Penman ADM and Charles J A (1973b). Effect of the position of
the core on the behaviour of two rockfill dams. Transactions of
11th International Congress on Large Dams, Madrid, vol 3, pp
315-339.
Penman ADM and Charles J A (1976). The quality and suitability
of rockfill used in dam construction. Transactions of 12th
International Congress on Large Dams, Mexico, vol 1, pp 533-556.
Penman, A.D.M. and Hussain, A. (1984). Deflection measurements of
the upstream asphaltic membrane of Marchlyn dam. Water Power
and Dam Construction, vol 36, no 9, September, pp 33-37.
Penman ADM and Charles J A (1985a). Behaviour of rockfill dam
with asphaltic membrane. Proceedings of 11th International
Conference on Soil Mechanics and Foundation Engineering, San
Francis~o, vol 4, pp 2011-2014.
Penman ADM and Charles J A (1985b). A comparison between
observed and predicted deformations of an embankment dam with
central asphaltic core. Transactions of 15th International
Congress on Large Dams, Lausanne, vol 1, pp 1373-1389.
Sowers G F, Williams R C and Wallace T S (1965). Compressibility
of broken rock and the settlement of rockfills. Proceedings of
6th International Conference on Soil Mechanics and Foundation
Engineering, Montreal, vol 2, pp 561-565.
Terzaghi K (1960). Discussion on Salt Springs and Lower Bear
River Dams. Transactions of American Society of Civil
Engineers, vol 125, part 2, pp 139-159.
CHAPTER 6
COLLAPSE: ITS IMPORTANCE, FUNDAMENTALS AND MODELLING
J . L. JUSTO

l. Introduction

The widespread construction of rockfill dams has its origin in the California gold
rush. Owing to the avai lability of rock and knowledge of blasting, the miners of the
California mounta ins used rockfiJI dams to store wa ter for dry season sluic ing of
place r deposits.
In many of t he eart ier American dams rock-filling was largely placed by cable-
ways and was sometimes rehandled by derricks. The rock was often dropped from a
considerable height, up to 9 m, with the object of breaking up the weaker rocks.
breaking off sharp poi nts and thin edges, and attaining, thus, better compaction . An
example of this is Strawberry Dam, 43 m high, and ended in 1916.
In some Algerian dams built in the 1930's, the whole of the rock fill was placed by
derricks or by hand, us ing specially selected rock near the upt ream face. T his
method has reduced porosity to 25-32%, and the success to avoid settlement will
be commented later.
T his technique gave place to that of "high lift" placing, in wich the embankmen t
was carried forward in successive end-tips. The rolli ng of the pieces of rock down
the end slopes was inte nted to have the same effect as the falling of rock as indi-
cated in the previous paragraph: The lift was from 7 to 66 m.
No special efforts were made to wet rockfill during placement.
Strawberry Dam was one of the higher rockfill dams at the time of its completion
(19 16). It was a concrete face dam (Gomez Navarro and Juan Aracil, 1958). The
wett ing du ring construct ion was scarce, but there was an important flood at this
period, that wetted the rockfill.
Dix Dam was, at the end of its constructi on (1925), a la nd-mark in rockfill em-
bankments. According to our records its height, 84 m, exceeded in more than 50%
the preceding heigh t reco rd. The rockfill was litlle slu iced with monitors during
placement, and a flood with a head of 18 m of water saturated the lower part of
97
E. Maranh., das Ne-.'es (<,d.). Advanct"s in Rockfill Slr"cl"re.f. 97-152.
C> 199 1 KI"""er Academic Publishers.
98

the rockfill during construction and produced a maximum settlement of 20 em due


to "collapse" of the rockfill (fig. 1).

NUMBER OF DAYS
5 10 15 20 25 30
w
(!)
r- ...J
...J
z -
~ ~ 0.25
ero
lJJ
n..I B

if)
h:
W
0.50
0

~~
~
0.75
-------~----~~~
w CROSS - SECTION
...J
~ OFDAM
w 1.0 0 L - -_ _---'-_ _ _ _L - -_ _---""~_"__"_-'>..L-_ ___'__ ___LJ

if)

Figure 1. Settlement of Dix River dam during a flood period (v. Nobari and Dun-
can, 1972).

Althoug the safety of the dam has never been impaired, leakage was, no doubt,
too large (2.7 m3 /s). It was constructed with up to 37 m lifts.
Salt Springs concrete face rockfill dam, 100 m high, ended in 1931, was the foll-
owing landmark. Sluicing ranged from 1/4 the volume of rockfill in the lower half
of the dam to much more in the upper part, with an average between the volume of
water and volume of rockfill of 2. This sluicing was estimated insufficient by Steele
and Cooke (1960). Leakage was 570 lis.
The same technique was used at Paradela (1958), 110 m high, with lifts of 66 m
and a volume of water of 4 times the volume of rockfill. Leakage was 3 m3 /s.
The importance of sluicing was clearly revealed during the construction of S. Gabriel
nQ 2 (Cogswell dam), concrete face and 85 m high.
The rock was cubic gneiss, with a compressive strength of 45 MPa, with less than
3% dust. It was placed in lifts less than 7.5 m high, and non sluiced.
"The contractor chose the use of shotcrete (gunite) as an alternative to concrete
for the sub-slab and the laminated facing. Construction of the facing followed pla-
cing of the rockfill and the packed rock part thereof, with a minimum of necessary
lag and by December 1933 had been about 33% completed.
On December 31, a major storm swept in from the Pacific Ocean which by noon
Jan'Jary 1, 1934 had yielded 382.8 mm of rain at the dam. The application of this
natural lubrication to the dry rockfill caused the latter to settle, specially that part
99

dumped in a gap at the right abutment. Inmediate settlement amounted to 4%.


Vertical settlement was accompanied by slight bulging in the upstream lower half
of sub-slab, particularly near the abutments.
Holes were drilled through facing and sub-slab and clear water pumped into
rockfill to hasten final settlement. After several months of this sluicing operation,
the afore-mentioned maximum settlement of 4 % had increased to 6 % (Baumann,
1960).
The damaged facing was replaced by a temporary timber facing. The leakage
when the dam was filled for the first time amounted to 3.7 m3 /s.
According to calculations, the porosity of the loose rockfill could be 40%, that
reduced to 26.7% during construction and to 22.5% subsequent thereto.
Probably the experience of this accident, and surely the performance of Salt
Springs dam were a help during the design of the two Lower Bear River Dams, 67
and 42 m high respectively. They were constructed in high lifts, and the rockfill was
well sluiced with high pressure monitors and a ratio between the volumes of water
and rockfill of 3. Leakage was 57 lis in Lower Bear No. 1. Dam No.2 suffered no
damage.
One of the former compacted rockfill dams is Schwammenauel Dam (Germany),
57 m high and built in 1938. Rockfill was compacted, in 1 to 1.5 m layers, by a fa-
lling plate 2.5 t in weight.
Compaction of rockfill spreads in Europe from 1950 (Sweden, Germany) and
Algier (1954). In the U.S.A. compaction of rockfill spreads from 1957 (v. Justo,
1968).

2. The fundamentals of collapse in rockfill

Collapse is the settlement of a soil, under load, when wetted.


There is a factor that has an extraordinary influence in the postconstructive de-
formations of a dam, and that is whether the rockfill has been wetted or not during
placement. A proof of that is what happened in San Gabriel nQ 2 dam.
This point has been checked plenty of times: if we load a sample of rockfill in the
oedometer (fig. 2), wait till the moment when settlements have practically ceased,
and later on, flood the sample, the rockfill will suffer very important settlements
produced by flooding (Justo, 1968).
This fact has been attributed to the lubricating action of water that produces new
slides inside the rockfill mass in search of more stable positions. This point of view
is held in the description of the settlements at San Gabriel nQ 2 made by Baumann.
100

In an oedometer water jets were prepared that allowed to inject water under pres-
sure at a given moment of the test. Both the injection of water under pressure and
under no pressure produced the same effect: an aditional settlement around 20%.
Even more: when water under pressure was injected to the saturated rockfill, no
additional settlement was produced.
Martin (1970) has carried out very interesting tests, wetting rockfill with water and
several organic liquids.
Table I shows several physico-chemical properties of the liquids and the settle-
ments of two rockfills when wetted by them.
It is interesting to notice that in water most of the settlement occurs in the first
day, but in the other liquids the settlement is very slow.
Cohesion is, finally, the result of chemical bindings, who are of electrical nature. If
a liquid introduces itself in a fissure, the dielectric constant or dipolar moment
must playa role in the loss of cohesion.
Of course other factors must intervene: surface tension can measure the pene-
trability of the liquid, viscosity governs the speed of penetration, etc.
It seems that the constant best related to collapse, for a high value of dielectric
constant, is the ratio between dielectric constant and dipolar moment.
This is an important research line, which has not been duly investigated.
According to Terzaghi the water needed is that necessary to saturate the rock. In
San Gabriel 2, the rainfall that produced a settlement of 4.5% was an 11% of the
rockfill volume. According to Terzaghi less than 20% is needed. Experiments quo-
ted by Veiga Pinto (1983) indicate that a 30% of water produces the same effect as
saturation.
There are other reasons for sluicing under pressure as it is to decrease segrega-
tion. This method increased the density of the compacted layers at Mont Cenis
Dam test fill.

3. Collapse modelling

Nobari and Duncan (1972) in an extensive work analyze the collapse of rockfill
when flooded. After wetting, the stress-strain curve of rockfill is the one corres-
ponding to the saturated material as indicated in figures 3, 4 and 5.
The apply their findings to the calculation of the stresses and strains of a dam with
a central earth core during filling of the reservoir.
101

~
2.0
0

, Satu roted
z 2.5 Ih.
'fO----<~-.:.. E1 Infiernillo Diorite
0
(J)
(J)
3.0 Max. Size 7.6 em
W
Q: (f= 1600 kPo
L. 3.5
:::!!
0
u 4.0
0 1 2 3 4 5 6 7 8 9 10
ELAPSED TIME - THOUSA NOS OF MINUTES

Figure 2. Collapse in rockfill during oedometer test (Marsal et aI., 1965).

Notwithstanding, the action of water on polished surfaces of minerals with a


three-dimensional crystalline lattice (the most frequent for rockfill employed in
dams) is antilubricating (v. Justo, 1968).
But the surface of rockfill is usually rough, and in this case, no doubt, water does
not modify friction at the surface (Justo, 1968).
The action of water in minerals of two-dimensional crystal lattice (mica, serpen-
tine, esteatite, talc, clorite, flogopite, etc) does produce a lubricanting effect, and
this might apply to slate or schist rockfill.
According to Terzaghi (1960), the larger part of the settlement produced by the
flood of pieces of rock with three-dimensional lattice (quartz, feldspar, calcite, etc.)
must be attributed to the loss of strength of the rock when wetted (mainly in its
outer surface), which produces the rupture of angular points and the subsequent
settlement of rockfill. The more weathered the rock, the larger is this lack of
strength. The rupture in contact points produces both: normal motions and slides,
as shown in tests carried out at Laboratorio de Geotecnia (Cedex, Spain).
Some argue against this reasoning that in the tests carried out by Sowers on the
Nantahala sandstone, there was no loss of compressive strengths upon saturation,
but there was an additional settlement when the rockfill was wetted at the labora-
tory (around 19% of the settlement of the dry rockfill).
Two comments to these tests can be made: first of all this is a very unusual case;
secondly the compressive strength depends upon the resistance of the whole mass
of rock and collapse depends upon the loss of strength of the outer, weaker parts of
the rock.
It has been said that the main advantage of sluicing is that water at high velocity
drags the fines towards the inside of rockfill. This is the philosophy that led Steele
and Cooke towards the use of sluicing with large amounts of water at high pressure
in Lower Bear dams.
The experiments by Sowers seem to reject this theory:
102

Table I. Collapse of marble fine gravel (2-5 mm) obtained by artificial grinding
when flooded with different liquids.
Collapse in 0.01 rrrn
Dielectric Dipolar Ratio Viscosity Surface
Liquid constant moment centi- tension Days
debies poises
E lJ E/lJ ruN/rn 1 2 6 25 50

lIater 30 1.9 43 1.0 73 128 133 137 -


(171 )
Formamide 109 3.3 34 3.3 58 30 42 50 86 103

Metanol 32 1.7 19 0.6 23 32 40 50 78 82


(61)
Etanol 24 1.7 15 1.2 23 14 25 52 76 -
(20)
Butanol 18 1.7 11 3.0 25 34 - - -
(27)
Nitrobencene 34 4.3 8 2.0 44 25 - - -
Bencene 2.3 0 0.7 29 12 31 31 - -
(32-61)
Kerosene 1.9 a - 2.0 28 7 - -
( 17)
Xi lene 2.3 0 0.7 30 23 - - -
Et i leneg l i col 41 2.0 20 20 48 22 - - -

(171) Collapse for a platy hmestone


Physico-chemical properties al 20 C
103

Water added

~-\
~ 2
~
~
;
I
+
.-0
c
dry
'-
.- 3
<II

.-C
<t 4
)(

Initially dry
5 then water added

200 400 600 800 1000 1200 1400 1600


Axial pressure (kPa)

Figure 3. Compression and collapse of Pyramid material in oedometer test (Nobari


and Duncan, 1972).
Cu = 14 5% < sieve 200 (ASTM) < sieve No.4 (ASTM)
I D = 70% 90% modified Proctor

Figure 6 shows the four effects of filling on the dam:


1. Collapse produced by wetting of rockfill
2. Water load on the foundation
3. Buoyancy in the upstream shell
4. Water load on earth core.
The order at which these effects occur is very important.
For layer A (fig. 6.1) which has just been saturated, the first effect is collapse un-
der the weight of the superimposed rockfill with its compaction unit weight. The
other three effects are of second order, for the layer affected, at this stage of the
water rise.
104

o ~ __________________________________________________ ~

Water added

~ 2
+-
II>
D ()-..,,......:a.-::-_ _+ _______~"_I_a= 0.53
o ....
)( 3 ............. o. Pc C
<! Measured ....
relax.ation ~............... Pc

4 ~---1~0-0----20~0--~3~0~0~~4~0~0--~50~0~~6~0~0~-=70~O~--8~O~0~--~
Ax.ial pressure ( k Po)

Figure 4. Axial stress relaxation for wetting without volume change for Pyramid
material (Nobari and Duncan, 1972).
< sieve No.8 (ASTM)
> sieve No. 16 (ASTM)
ID = 83%

We shall refer mainly to collapse. To understand the modelling of collapse it is


useful to refer to the simple, one-dimensional rheological model of figure 7. The
rockfill in the upstream shell is subject to a pressure, p A, resultant of the weight of
the material placed above during construction, PW, and the vertical constraints
applied by others elements, PA - Pw> represented by the spring. When the shell is
flooded, collapse of the soil will produce a deflection of the top of the specimen,
and, this will lead to a reduction of the total pressure in the specimen. The stresses
and strains are shown in the lower part of figure 7, in which the initial (prewetting)
condition is represented by pointA, and the final equilibrium condition by point C.
An analysis of the effects of wetting could be performed in two steps as follows:
First let us assume thaI we, somehow, hold the plate in its position. Then, the path
followed would be AB, and in the FE method would be equivalent to a reduction of
stresses, to adapt the strain state of each element, that up to now has remainedstiII,
to its new stress-strain characteristics.
105

1400

1300 ........ ----- ...............


Dry".."
1200 -~-/ ..;~ ....

""
-- --
1100 /
0
a..
.:Jt:
1000 /
/ " -----
Water I ,. .... '
b'" Added I
I 900 I
./ ,. '" "", Wet
b-
- ~/ " "
""
CI)
800
I
"
CI)
lLJ
0: 700 '/
I- ~
CI)
600
0:
0
I-
c::{
500
>
lLJ 400
0

300

200

100

4 6

--
Z 8 10 12 14 16 18 20 22 24
::il-
~z 0
Cl)lLJ
t.)t.)
_0: 2 - __
-------- _-------- --h- Dry
O:lLJ
t;ja.. 4
Dry -Wet Wet
::!E I
~~~~~-~-~-~~
~
...JIJI
0
>
6
------ I

> 0 2 4 6 8 14 10 12 16 18 20 22 24
AXIAL STRAIN, a (%)

Figure 5. Comparison of stress-strain and volume change curves for dry, wet and
dry-wet specimens in triaxial compression for Pyramid material (Nobari and Dun-
can, 1972)
0'3 = 300 kPa < sieve No.8 (ASTM)
> sieve No. 16 (ASTM) ID = 90%
106

Collapse Due
to Wetting

(1) Collapse due to wetting Upstream shell material

(2) Water -load on foundation

(3) Buoyant uplift on upstream shell

(4) Water-load on core

Figure 6. Effects of reservoir filling on a zoned dam.


107

Rigid to
Fixed
boundaries
Compression
ring

c: A Dry compression
_1
-- --
o
....en
'- cr--.....a--~
Relaxation at--
constant volu me

recompression ...... Wet compression


'< ....

Axial pressure, p

Figure 7. Schematic illustration of simultaneous stress and strain change due to


wetting in one dimensional compression apparatus.

But in the rheological model, the decrease of stresses in the compressed soil leads
to a lack of equilibrium in the top plate between the stress in the upper part of the
plate, PA, and the pressure in the rockfill, PB. This difference will act when we free
the plate, compressing the sample that now obeys a different stress-strain law (path
Be in figure 7). This second step is carried out in the FE method transforming the
difference between the initial stresses of the element (PAin figure 7) and the new
ones (PB) corresponding to the initial strain under its new stress-strain law in nodal
forces, which now will act on the dam body.
This method is valid for any constitutive law.
Justo and Saura have developped a three-dimensional FE method, linear-elastic
108

in principle, but considering the no-tension strength of rockfiII, and different mo-
duli according to the direction of stress-change. In our case collapse has been simu-
lated by a decrease in the modulus of elasticity of the material, with the following
steps:
1. Relaxation of stresses in submerged elements, assuming there is no change in
displacements, in a proportion a, that is an entry of the program, obtained from
collapse tests.
The resulting stresses in the elements are (I-a) uij' and the residual stresses a. Uij
2. The residual stresses a.u ij are transformed into nodal forces through the corres-
ponding B matrix at each element, and the set of nodal forces of all relaxed ele-
ments is introduced as a load external to the structure.
3. The modulus of elasticity of the submerged elements becomes (I-a) E, where E
is the modulus corresponding to the as compacted rockfiII, obtained from measu-
rements during the construction period.
Hooke's law before and after relaxation is indicated below:
Before relaxation:
1
Eii = ---[uii-v(Ujj- ukk)] (1)
E

1
IJ
Y ij = ----- (2)
G
i = x,y,z
j = x,y,z
k = x,y,z
iT j ~ k

After relaxation:
1
Eii = --------- [0 jj (l-a)- v [0jj (1-a)- 0 kk (1-a)] (3)
E(1-a)

1 ij(1-a)
Y ij = ----------- (4)
G(1-a)

So, we see that relaxation is equivalent to a reduction ofE up toE(l-a}.


Measurements of lateral stresses carried out by Valstad and Strm (1975) and
Veiga Pinto (1983) indicate that there is no change of vdue to submergence.
Veiga Pinto (1983) finds that v decreases with stress level for oedometric condi-
109

tions. For 0'1 from 500 to 1600 kPa an acceptable average value of \! is 0.22.
4. The stresses obtained from the action of these nodal forces must be added to
those existing before filling of the reservoir. In the submerged elements the cons-
truction stresses have been corrected multiplying by a factor (i-a).
A similar approach, but ussing a variable, hyperbolic, E value during construc-
tion is used by Alberro et aI. (1976), and Veiga Pinto (1983).
Naylor et aI. (1989) present the application of the same general method to any
constitutive law, and find with the K-G method, for Beliche dam, rather different
results than Veiga Pinto (1983) with a hyperbolic method. It will be interesting to
compare both calculations with the measured displacements during reservoir filling
to find the applicability of different constitutive laws.
Tables II and III show the variation of a in different types of rock with compac-
tion and stress level. Of course if the stress-strain curves are linear, then a is a
constant. It is possible to introduce different a values in the elements depending
upon stress level. At least in some materials the election of an average a value is
not more inaccurate than to obtain a from laboratory tests.
There is often (although not always) an increase of a with stress level. The infl-
uence of compaction is not clear. The relaxation coefficient is larger in soft rock.
Our FE method has been applied to Infiernillo dam. Values of a ranging from 20
to 30% have been chosen based upon laboratory tests and average stress level
(Justo and Saura, 1983). This allowed us to study the influence of a in the results.
Figure 8 shows the layout of instrumentation. Figure 9 shows the FE discretiza
tion.
Buoyancy in submerged rockfill has been taken into account generating nodal
forces equivalent to the corresponding uplift, and water pressure on the core is
introduced as external nodal forces on the upstream face of the core (v. Justo et
aI., 1989).
For buoyancy and water pressure two hypotheses concerning the modulus of de-
formation of the upstream shell have been made:
a) Maintaining the same modulus (Eo) obtained during construction.
b) Introducing a modulus much higher (E= 4 Eo) to take into account that they
are processes of unloading.
Justo (1990) has measured moduli during unloading ranging from 2 to 11 times
the values for loading in low grade rockfilI. As Infiernillo rockfill is sound grano-
diorite, we thought, that 4 might be a good average.
Figure 10 collects the calculated deformations in the vertical axis of the dam when
it is subject, separately, to the following processes (for E/Eo = 4):
- collapse
- buoyancy
- buoyancy + water loading.
The graphs indicate the small influence of buoyancy in the deflections of the dam.
110

Table II. Relaxation coefficient, a, as a function of stress level from oedometer test.
Stress
Type of rock COIIl>action level a Reference
(kPa)

Mar 1y 1irrJestone n=38.2% 200 0.26 FlITIBga II i , 1969


700 0.25
200 0.30
600 0.30
average 0.28
Argillite 10=0.7 200 0.3 Fig. 3
1600 0.5
10=0.93 200 0.33 Nobari and Duncan, 1972
1600 0.49
10=0.83 200 0.36
1600 0.58
oven-dry 100 0.53
400 0.69
10=0.83 400 0.45 Fig. 4
800 0.53 relaxation test
average 0.48
Tonalite n=36% 200 0.60'0.35 Bertacchi & Bellotti, 1970
3000 0.190.21

average 0.34
Serpentine n=39.5% 200 0.34'0.30
3000 0.28'0.31
average 0.31
~eathered schist 10=100% 400 0.57 Veiga Pinto, 1983
and graywacke 2000 0.53
average 0.55
III

Table III. Relaxation coefficient in crushed basalt > 0.59 mm (Donaghe & Cohen,
1978).

Upper Stress a
size n level
(11111) (kPa) (%)

50.8 0.27 ~800 0

5500 0.30
25.4 0.29 gOO 0
5500 0.17

12.7 0.32 5500 0.43

6.35 0.36 ~400 0

5500 0.39
76.2 0.28 ~200 0

5500 0.35
50.8 0.30 ~200 0

5500 0.30
25.4 0.32 ~400 0
5500 0.32
12.7 0.36 200 0.49

5500 0.35
6.35 0.40 200 0.49

5500 0.39

Average 0.23
112

St.O+OO _ . _"'-.-=r>~_ 160


140
--::=:::::::::::::::::~12~0 100

.. B

160
St. 0 + :3 50 ~~.~t:===::::::----------- 180 200
o 50 100 200
r---=t I I
PLAN VIEW
Graphic scale (m) LEG END

-$- VERTICAL CROSS-ARM (D)


INCLINOMETER (I)
Il SURFACE REFERENCE

Crest. EI. 180 POINT


180.--~~~~~----~~--~~~
1601------~~~~~~
140~==~~~~~~
120 r----=-:--:-=-::c-=----;?'"
100~----~~
80'--------./ EI. 80
Zone 5

SECTION B, ST. 0+135

I 80 I-I I-II I-V !-IX


I Iii l'
HiO .. ~ j,f'
140 ",\.. ,/J"
120 "'%, ~..
I~g .,.,~ . 1/
60 .. ~ v-
. .... A:'
~g "left abutment .. ;~-. Righ abutment-
o 50 100 200
SECTION A r-----+-J! 1
Graphic scale (m)

Figure 8. Infiernillo darn and location of instruments.


113

CREST LEVEL 180


MAX W. L. 165 -I=: Z
\I AXIS OF SYMMETRY OF DAM
2 nd STEP
W.L.120 ~
'\
\I ~~
1 sf STEP~~
ty:<Q

x
DUMPED TRANSITION
ROCKFILL

ACTUAL LINE OF DIVISION BETWEEN


BOTH KINDS OF ROCKILL

Figure 9. Discretization of Infiernillo dam in finite elements.


114

x- AXIS DIRECTION Z - AXI S DIRECTION

I
I
140
II I.

I 120
l\
:\
I'

\ 100

\
I .

\
I
I
80 I
\ I

\
I
I \

" 14<;/./
I
I
/ I

"."
,/

,-'
f
.,-'
-20 -10 0 10 20 em -20 -10 a 10 em
UPSTREAM DOWNSTREAM SETTLEMENT HEAVE

COLLAPSE

BUOYANCY + WATER LOAD


BUOYANC Y

Figure 10. Displacements at the axis of the dam during filling of the reservoir up to
normal top water level (Justo et aI., 1989).

In June 1964, the dam was filled up to a height of 90 m, when a flood overflowed
the cofferdam.
This situation was simulated by us and the calculated results are included in figu-
res 11 to 13 together with the displacements measured at the inlinometers I-I, I-II
and I-V (v. fig. 8)
These figures indicate a substantial improvement in agreement between calcu-
lated and measured results when a larger stiffness is assumed for the upstream
rockfill in processes of unloading (E/Eo = 4).
The model reflects rather precisely the upstream tilting that showed the core
115

when the reservoir had arrived to this level.


X-AXIS DIRECTION Z-AXIS DIRECTION V-AXIS DIRECTION

140 140

120 120

100 100

80 ,. 80
I
I
60

40

20 20 20

~ -- -- T- - , - - ,-,----, - ,- 1 - ~I- ~ ,,-.-+-.---.-


-10 10 20 em -10 5 em -10 Sem
UPSTREAM DOWNSTREAM SETTLEMENT HEAVE DAM CENTER ABUTMENT

CALCULATED WITH E/Eo=4


_ . - CALCULATED WITH EI Eo =I
______ MEASURED

Figure 11. Displacements at inclinometer I-I for a water height of 90 m (Justo et aI.,
1989).

X- AXIS DIRECTION Z- AXIS DIRECTION Y-AXIS DIRECTION

/
120
120
/ 120

E
100
/
/ I-
:I:
100 100
/
i
C>

/ w
:r:

I
80 80 80
/
60
0 2 4em -2 0 2em - 2 0 2 em
DOWNSTREAM SETTLEMENT HEAVE DAM CENTER ABUTMENT

CALCULATED WITH E I Eo= 4

_-0- CALCULATED WITH EI Eo= I


- - - - - - MEASURED

Figure 12. Displacements at inclinometer I-II for a water height of 90 m (Justo et


aI., 1989).
116

X - AXIS DIRECTION Z-AXIS DIRECTION Y -AXIS DIRECTION

, 140 140

r 120 120

r
100
\\, 100

I
80 80,
, "-
/
- 60 - 60
E
~
I
(!)
40 40
t 40

t
w

f 20
I
20
20

-10 10 em -10 10em -10 10 em


UPSTREAM DOWNSTREAM SETTLEMENT HEAVE DAM CENTER ABUTMENT

CA LCULATED WITH E / Elf 4


- - . - CALCULATED WITH E / Eo= I
- - - - - - MEA SU RED

Figure 13. Displacements at inclinometer I-V for a water height of 90 m (Justo et


aI., 1989).

We have also studied the situation of the dam when the normal top water level
was reached. Figures 14 to 16 refer to this situation.
Figure 15 collects the displacements at the vertical of the inclinometer I-I where a
direct comparison between measured and calculated displacements can be made.
In this figure a substantial improvement in the agreement between measured and
calculated displacements is reached when E/Eo = 4; The transverse movements
calculated under the hypothesis E/Eo = 1 are nearly four times the measured va-
lues.
117

X- AXIS DIRECTION Z- AXIS DIRECTION

~ .'-...,. 140

120
"'" '" 120

lOa
"'- 100
t-
:r:
\
\
(!)
80
w

I
:r:

60

--. --. _.I


40
-. --
. .

--. --.
""'-
20
"-
:::-- .

10 20 30 40 50 em -10 10 em
DOWNSTREAM SETTLEMENT HEAVE

CALCULATED WITH EI Eo= 4


CALCULATED WITH E/Eo= I

Figure 14. Displacements at the axis of symmetry of the dam during filling of the
reservoir up to normal top water level (Justo et aI., 1989).

The vertical displacements calculated under both hypotheses are similar, specia-
lly in the lower third of the dam.
Figure 17 compares the measured downstream movements at Oroville's dam and
other displacements calculated by Nobari and Duncan using a plane-strain FE
method based upon the hyperbolic stress-strain relationships developed by Kond-
ner and Duncan.
We see that the measured movements are smaller than the calculated ones.
The difference between the stresses measured by different methods in this dam
reflected the difficulty of properly measuring these otresses, and the only agree-
ment with the calculated stresses was that they were both quite small.
118

X-AXIS DIRECTION z- AXIS DIRECTION Y- AXIS DIRECTION

140

120

100

80

50

40

20 20 20

20 40 50em -20 -IOem -10 10 em

DOWNSTREAM SETTLEMENT DAM CENTER ABUTMENT

CALCU LATED WITH EI Eo=4


CALCU LATE D WITH E I Eo= I
MEASURED

Figure 15. Displacements at inclinometer I-I for normal top water level (Justo et
al., 1989).

Alberro et al. (1976) have also used a plane-strain hyperbolic FE method for the
study of the filling of Jose M~ Morelos rockfill dam with poor IO;!SUItS.
119

X -AXIS DIRECTION z- AXIS DIRECTION Y-AXIS DIRECTION

'-....
'-..
I
I
I \ ,
140 ' \ \ 140
I
'\
............
"
I
" 120
\\ 120
120
''\..'" \\
\'

) \
, \ ,

100
/
100
\\ 100

,
,-' -,-'-,-'
I
"
80 / 80 I 80

60 60
t-
:x:
(!)

w 40 40 40
:x:

20 20 20

10 20 em -20 -10 em -10 5em


DOWNSTREAM SETTLE MENT DAM CENTER ABUTMENT

CAL CULATED WITH E 1 E 0=4


CALCULATED WITH EI Eo= I
MEASURED

Figure 16, Displacements at inclinometer I-V for normal top water level (Justo et
al., 1989),
120

Crest EI. 922

Instrument
House T

5 4 3 2 Instrument
House U
EI. 165
Graphic scale (m)
4 3 2
I I
o 50 100 EI.IOB

Location of horizontal movement devices

E 15
0

-
c:
0
Co Icu lated
0 10
CI)

"-


CI)


-0

E 5
-
0
....
CI)

U)
c: 0
~
0 7 6 54 3 2
0 Location
Horizontal deflections at EI. 165

10
E
E~ ~ Calculatec

'M"'~
c

--
c:
Q.l
.2 5
.'
L-

U)
c:
0
CI)
~:;::::
-0-....
-0-
--0
o Q.l
0-0 0
7 6 5 4 3 2 U
Location
Horizontal deflections at EI. lOB

Figure 17. Comparison of calculated and measured downstream deflections in Oro-


ville darn caused by reservoir filling (Novari and Duncan, 1972)
121

4. One-dimensional collapse during water rise in a granular material. Buoyancy


and creep.

So as to help in the understanding of the phenomenon of collapse, and the sequ-


ence of occurrence of collapse and buoyancy we shall study the collapse during
water rise in a granular material (fig. 18).
Figure 18a represents the situation at the end of construction:

o = Overburden pressure at end of construction at the height of final water rise


H2
Eoe~ = oedometric modulus of compacted material (for loading).
y = unit weight of compacted material below H 2 .
Figure 18b represents the water rise:
E~ed = oedometric modulus of saturated material (for loading).
E~~ = oedometric modulus of saturated material for unloading.
y' = unit weight of submerged material below H 2 .
Figure 18c represents the stress strain path for a layer at level z. Point A repre-
sents the situation at the end of construction, point B when water reaches level z,
and point C when water reaches level H2.

The collapse settlement of the layer affected by a rise of the water level, dz, will
be:
1 1
ds c = [o+y(H2-z)](---~--- - --d---) dz (5)
Eoed Eoed

The collapse settlement of level H, when water rises from level H1, below, to level
H 2, above, is:
1 1 JH
sc = (---5--- ----CC-) [o+J( H2-z )]dz (6)
Eoed Eoed HI

1 1
sc = (---~--- - ---ci---)[(o+(H2)(H-H l )- V2y(H2-H2 1)] (7)
Eoed Eoed

1 1
sc = ---d---)[( o-t'yH2)- V2y(H-Hl)](H-H 1)
(---~--- - (8)
Eoed Eoed
122

1 1
Sc = (---5-- - ---a---)[o+yzY (H 2-H 1 +H2-H)](H-H1) (9)
Eoed Eoed

f ina I wate r level


z + dz
z

HI

(a) ( b)

I
0-+6' (H2-z) (J'tt(H 2-Z)

o t t t> stress

construction
slope 1/ Eged
stress relaxation
c

-
c straining under residual stress
~

( /) s lop e 1/ E~ed
collapse

buoyancy
slope 1/ E~ed

(c )

Figure 18. One-dimensional problem of water-rise in a granular material.


a) End of construction
b) Water rise form level z to level (z+dz)
c) Stress-strain paths in layer at level z.
123

If we call:

the stress at the middle of the layer form HI to H at the end of construction, then
equation 9 becomes:

1 1
Sc = (---~--- - ---;3"--) av(H-Hl) (10)
Eoed Eoed

The collapse settlement of a layer of thickness H-H I is equal to the difference bet-
ween the unit strains, produced by the compaction unit weight, at the middle of the
layer, for saturated and compaction conditions, multiplied by the thickness of the la-
yer.

If HI = 0, then:
1 1
Sc = (---5--- - ---ci---)( + Y H2 - VzYH)H (11)
Eoed Eoed

1 1
Sc = (------- - --d---)OavH (12)
E~ed Eoed
When all collapse has taken place in a layer, buoyancy is still infinitesimal.
When water reaches the level H2, buoyancy at level z produces a decrease in ef-
fective stress of value:

6.0' = 0 + y'(H 2 -z)-0-y(H2-z) = -( y-y')(H2-z) (13)

The extension produced by this decrease in stress in the layer of thickness dz is:
(y-y')(H2-z)
dSb = - ---------------- dz (14)
su
Eoed

The heave at level H will be:


(y\,,') H
sb = - -------- I(H2- z)dz (15)
Eg~d
124

(16)

The buoyant heave of a layer of thickness H is equal to the uplift strain at the centre
of the layer for unloading conditions, multiplied by H.
If there is an initial water level HI, to find the buoyant heave at H for a rise from
HI to H2' it is necessary to substract the heave corresponding to the rise up to Ht
y-y'
sb 1 = - ------- 1/2H 21 (17)
su
Eoed

Substracting (17) from (16):


y -y'
sb12 = - --su-- (H2H - V2 H2_V2H21 ) (18)
Eoed

An application of these equations will be made in the Appendix.


Respect to Nobari and Duncan's approach collapse of the layer at level z will be
modelled in the following way:
1. Stress relaxation of value a.[ o+y (H2-z)] under conditions of no strain-change.
The resulting stress will be (I-a)/o+y (H2-z)] (path AD in figure 18 c).
2. Straining of the layer of saturated material under the residual stress a.[o+y (H2-z)]
(path DB in figure 18c).
3. The residual stress is added to the relaxed stress to find the actual stress:
(I-a)[0+Y(H2-z)]+a[0+Y(H2-z)] = o+y(HXz)
Buoyancy follows in the same way indicated previously.
At least in this one-dimensional problem, the asumption that collapse occurs si-
multaneously in the whole layer H, followed by buoyancy, leads to the same result
than modelling collapse followed by buoyancy in each layer, because buoyancy only
affects the layers below water level, and collapse is produced by the weight of the
layers above.
When the water level rise is slow, creep may be superimposed to collapse.
Creep is sometimes represented by the action of a fraction of unit weight, fy(v.
Segovia, 1989), where f is the "creep coefficient".
This is equivalent to a decrease of modulus of deformation, with time, from E to
Eo where:
fY y y
125

f 1 1

E
Ec = -------- (19)
1+ f

In figure 19, creep of the layer at level z > Hi would be:

0+ Y(H2-z)
dS cr = ----cr------- f dz (20)
Eoed

- - - - - - - z.+dz
_______ z

Figure 19. One dimensional creep in a granular material.

o+y(H2 -H 1) +y'(Hrz)
ds cr = ---------------------------------
s f dz (21)
Eoed

The creep settlement at level H would be:

I HI I H
ser = f[---s---I[o+y(H2-H 1)+y'(Hl-Z )]dz+ ---Cf--I[o+y(H2-z)]dz]
Eoed 0 Eoed HI
126

f
scr = ---~-- ([CH'1 (H Z-H 1)+'1'H 1]H 1 - 1/2 '1 'HIZ)+
Eoed

f
+ ------- ([o+'1HZ][H-HI]- Vz '1 [HZ-HIZ])
~ed
f
scr = ------- [0+'1 (Hz-HI) + Yz'1'H I ]H 1 +
E5ed

f
+ -------- [ 0 + '1 (HZ-H) + Y2'1 (H-H 1)](H-H 1) (ZZ)
Eged

o'avl av2
s = f[------- HI + ------- (H-HI)] (23)
cr s d
Eoed Eoed
where:

0' avI = average effective stress at level Yz HI

oav2 = average stress at level Yz (H + HI)

(24)

E av1 = average strain of layer HI

E av2 = average strain of layer (H-HI)

The creep settlement is equal to the sum of the average strains, produced by gravity
forces, of the different layers, multiplied by the thicknesses of the layers and by the
creep coefficient f.
For a homogeneous embankment we substitute in equation 22:
0=0
H2 = H
HI = 0
H = z
127

and have:
fY 1
scr = ---a---
(H - ---- z)z (25)
Eoed 2

5. Collapse produced cracks

Collapse of upstream shell has produced longitudinal cracks, usually in the contact
between upstream shell and core, in many earth core dams constructed with dum-
ped rockfill, specially when the inclination of the core is small:
Bear Creek (65 m high), Cedar Cliff (50 m), Cherry Valley (100 m), etc.
Longitudinal cracks have also appeared in clay core dams with compacted rock-
fill:
Cougar (136 m high; v. fig. 20); Djatiluhur (100 m high), with cracks penetrating the
core; Gepatsch (130 m).
Some have also suffered transverse cracks in the core: Mattmark (120), Cougar,
and Infiernillo (148 m high).
Figure 20 shows the relationship between pool elevation, settlement and downs-
tream movement at Cougar dam. It is easy to see that there is an increase in the
rate of settlement when the water rises, and some elastic rebound in the downs-
tream deflection for water lowering, but much smaller than the primary downs-
tream deflection, showing that the loading modulus is many times larger than the
unloading and reloading ones. The upstream shoulder settles more and deflects less
towards downstream than the downstream one, due to collapse, and the appearan-
ce of three longitudinal cracks.
In Infiernillo (fig. 21) transverse cracking of the core, as a result of the sudden fi-
lling of 90 m of water, appeared when water was at level 120, and the measured
tension strains approached -0.1 % near the abutment.
At the end of construction, no tension stresses appear in the calculation, in the last
layer, which agrees with the absence of transverse cracks in the core at that stage.
With water at level 120 comulative values of longitudinal strain and stress of
-0.51% (extension) and -2.7 kPa (tension) have been obtained. The cracking of the
core coincides with the appearance of tension stresses, as indicated by other aut-
hors.
Figure 22 shows a cross-section of Canales rockfill dam, 158 m high, the highest
embankment dam in Spain.
Rockfill was compacted in 1.5 m layers.
First impounding produced a differential collapse settlement aroun 85 cm (fig. 23)
between the upstream shell and the core on one side and the downstream filter and
rockfill on the other side. A longitudinal crack was produced (fig. 24)
128

Overbuild

Typical Cross - Section

E
Diversion tunnel closed sept. 24, 1963 (pool elevation 396 m)
Q)
550
>
~
500
W+-W v ....,
0
0 ,.... V "-' rv~lllll.\
IV 'Pool Elevation ,..r--
i.-.r'
Il... 450
100
E Hubs 11 8. 12 set Dec.1963
~
+- 75 k pstream Shoulder
c f-'
Q)
50
E v-
V -
T St,une,
Q)

25
Do rnrela
+-
+- i--"
Q)
(/)

60 Downstream Shoul,oer
E c
g .!:!~ 40 ~
-t;tE20 f
c Q) U ./
3::;:- t- f - /
Upstream Shoulder
o Q)
0"0 1964 1965 1966
Crest Deformations and Reservoir Level
120 m of crack
90 m of crock 15 cm maximum vertical offset
30 cm max. vertical offset 76 m of crack
Hub 12 Hub "no vertical of fset
~;)==i=I'-r-

Differential settlement and crocks on crest

Figure 20. Crest movements and cracks in Cougar dam (Pope, 1967)
129

Right
abutment
Left
abutmet
~
Crest settle ments
~~~--------------------~~~O

",' . ,"
"'"
./
./
/
/:;;,
/
10

'x-.. '. 0 __ ~_.,,! 20


-... ....... ~)( ~--
Legend
*""*Observed,W.L.120m --.- - ..... - ---
"'.--::. -x-- -x- _ -x-.;::.7 30
40
*" *" II ,,165m
~Calculated .. 120, P = 0.2
o 0 " 165, "
......
-+-+-
"
' , 120, P =0.3
" 165, "
-a- Observed 1971

Upstream

30
20

10

, o
Left
abutment
.........
- ... _- ....--_.- ....../ o

Right
10

abutment 20

o 30
o o
Tranverse crest displacements
Downstream
during water rise

Compression
Right
abut ment 0.2
x
fine / -
/'\~- ..... - - ....... .........
Left era ck s I ~/~::::~t::3;?~~ O. 1
abu t men t---i I},-/_=--"'9'7"'~c-...{----"'----------'''''''''''=-''---=:------1r-i'-~i-=0
" ...., IX 0.1 '--~...rJ1

" I x

'.1
0.2
Longitudinal strains in '1 I
crest during water rise II 0.3
x

Ex te nsion
Figure 21. Displacements and strains in crest during water rise at Infiernillo dam.
130

13.00
+f- 965.50
!L..-

845.00
.'~ =-
.' .. :~o:.{) >'", " ~.~ "\q
: ..: , . t . -':".

10.00

925.00

1 st phose

Figure 22. Canales rockfill dam (Bravo et al., 1988)


a) Original cross-section
b) Modified crest to cope with tensile strains
1. Impervious core
2. Filter material
3. Rockfill
131

Figure 23. Canales roekfill dam. Differential settlement, about 85 em, between core
(to the right) and downstream filter and roekfill (to the left).
132

Figure 24. Canales rockfill dam. Differential settlement between core (to the left)
and downstream filter and rockfill (to the right). Observe longitudinal crack in con-
tact.

6. Collapse and post-constructive settlements of rockfill dams.

In this section several examples of collapse and post-constructive settlements in


dams will be presented.
It will be shown also how from preliminary measurements in dams the initial pa-
rameters for FE compactions can be obtained assuming one-dimensional condit-
ions.

6.1. CENTRAL CORE DAMS.

Postconstructive settlements may be related to creep or collapse.


In central core, aproximately symmetrical, dams, the difference between the sett-
lements of the upstream and downstream shells during filling of the reservoir is re-
lated to collapse.
133

Yeguas dam was ended at the beginning of December 1987. Figure 25 shows the
lines of equal settlement during construction.
The settlements in the upstream and downstream parts of the core are appro-
ximately equal.

______
-
Schist and
raywacke
&
rockfill
186.00

'"

--t-.JJ.'-----l-~~_t- -- . .-
small reservoir for
construction PlXPoses
SECTION 0+210

Figure 25. Yeguas dam. Lines of equal settlement (m) during construction (begi-
nning of December 1987).

On the other hand, the settlements in the upstream shell are larger than in the
downstream one, because, during construction, a small reservoir (fig. 25) was held
between the upstream cofferdam and the core for water needs during construction
(between levels 171 and 187).
Figure 25 shows the lines of equal postconstructive settlements when the reser-
voir rose to the level indicated in the figure at the end of June 1988.
In Appendix 1, with the settlements of cells 10 and 16 during construction and du-
ring water rise up to level 213.34, assuming oedometric conditions, the following
parameters of the conglomerate have been found:
Modulus of linear deformation of compacted material (loading):
Ed = 38 MPa

Idem for saturated material (loading):


ES = 25 MPa

Idem for saturated material (unloading):


ESU = 99 MPa
134

Poisson's ratio (assumed):


v = 0.35
Relaxation coefficient:
a = 0.35
Creep coefficient:
f = 0.115

These parameters might be used as preliminary parameters for a finite elements


calculation of the dam.

SECTION 0+210

Figure 26. Yeguas dam. Lines of equal postconstructive settlement (m) when the
reservoir reached the level indicated (end of June 1988).

Figure 27 shows the beautiful aspect of the dam in August 1988. No cracks have
appeared on it. Leakage ranges from 2 to 4l/s.
The conglomerate was compacted in 40 cm layers, with six passes of a vibrating
roller with a mass of 10 Mg.
135

Figure 27. Yeguas rockfill dam (Spain) in August 88. Designer Engineer A. Pas-
tor.

6.2. MARTIN GONZALO ROCKFILL DAM.

In Martin Gonzalo rockfill dam impounding started the 2nd of Ferbruary 1987.
Leakage ranged from 4.2 to 9.S lis till the 26th of November when the height of
water was 31.6 m (fig. 28).
The 27 th of November several ruptures in the membrane produced a sudden in-
crease of leakage up to 1000 lis (v. Justo, 1988).
The dam was emptied, and, after a provisional repair, the membrane was retired
and substituted by an asphaltic facing.
Figure 28 shows the height of water inside the dam during the accident.
A preliminary evaluation of the parameters of the dam, assuming one-dimensi-
onal conditions, has been made in Appendix 2, with the following results:
Modulus of linear deformation of compacted material (loading):

Ed = 11.7MPa

Idem for saturated material (loading):


136

ES = 10.0 MPa

1.5

i= 1.25 %
. .. :"~-.- ..
:~; '.:.',,~

o 10 20 m

Figure 28. MartIn Gonzalo rockfill dam. Water level and piezometric head the 1st
of December 1987. (Justo, 1988).

Idem for saturated material (unloading):

ESU = 235 MPa

Poisson's ratio (assumed):

v = 0.3
Relaxation coefficient:

a = 0.15

Creep coefficient:

f = 1.3

The creep rates included in figures Ap. 2.2 and Ap. 2.3 are much larger than mea-
sured in other dams (v. Parkin at Chapter 9 of this book).
We see in figure Ap. 2.4 that wetting of the lower part of the dam increases the
creep rate not only in the wetted part of rockfill, but mainly in the upper (non wet-
ted) part. This fact has also been observed in Yeguas dam (v. fig. 26).
137

6.3. POST-CONSTRUCTIVE SETTLEMENTS OF ROCKFILL DAMS

Post-constructive settlements may be related to creep or collapse.


Figure 29 shows the regression of the maximum crest settlement against dam
height for dams of dumped rockfill poorly sluiced.
Strawberry (St) and Dix (Di) dams, which were little sluiced during construction,
and also crossed by a flood have had smaller post-constructive settlement than ot-
her dams that have not been sluiced at all during construction, such as San Gabriel
(San 2).

6
San 2.
5
4
+M MUWNGUSHI (SCHIST) I
3 I i I I
E
(,)
I V
2 '7
I-

VIIe; ~i
Z
w M
::E +
w
-I
l-
I-
10 2
Sw 1/ I

W 8 /
C/l
/
I- lI.st
6
C/l
w /
ex:
u
5
4
80/
::E 'I S
::::> I I I

/
::E 3
F
x

::E 2
V 8'

10 2
10 2 3 4 5 6 8 10
HEIGHT OF DAM (m)

Figure 29. Regression of maximum crest settlement versus dam height. Upstream
face dams of dumped rockfill poorly sluiced (Justo, 1986).
138

Figure 30 shows the same graph for well sluiced dumped rockfill dams. We see
that there is no advantage for using a high amount of water, such as happened in
Paradela (Pa) compared with smaller quantities such as in Salt Springs (Salt), as
commented in section 1.

~Jji
I I
(~Qlt6
10 2
Coul
8
6 (Wi)"
E
u
5
r- 4
(L 0) II
z
w 3 II
:E
w
...J 2
(Wis) li ( Is)
r-
r-
w
~ II
(Vi)
CI)
I ) en
(L02)
I I
r- IO 1 - - ( B.il
CI) / (N z)
w 8
a:: I7:-{M
u
6
:E 5
J
:::>
:E 4 /
x
<:(
:E 3 /
J (C on)

IT')
2 )(

10 2 3 4 5 6 8 102 2
HEIGHT OF DAM (m)

Figure 30. Regression of maximum crest settlement versus dam height. Upstream
face dams of dumped rockfill well sluiced (Justo, 1986).
139

Figure 31 represents in the same graph the regression lines corresponding to four
ways of placing rockfill. The two non-sluiced types of rockfill dam (a and b) have
the higher settlement, and compaction reduces greatly the post-constructive settle-
ments.
<+
3

2
v
V
II
I
(b)
10 2 7
( c)
8 I
6 I
E
0
5 I
I- 4 I ~'
z
w 3 I 1I
::E
w
...J
l-
I-
2
[7
/ 17/
~!

/
W
C/)
kI/ II
---
/
I- I~ (d) ./
I
~ 10 7 7 7 \
a:
u 8 7
I
-, I
./
\/

::E 6 I V
::>
5 I I I-
::E
x 4
V
/ / v
<t
::E
3
.I II ;"

./
/7 / //

J./
2

.///
10 2 3 4 5 6 8 10 2 2 3
HEIGHT OF DAM (m)

Figure 31. Regression lines corresponding to upstream facing dam, and four ways
of placing rockfill:
a) Hand or derrick placed
b) Dumped, poorly sluiced
c) Dumped, well sluiced
d) Compacted, three-dimensional crystalline lattice.
140

7. Conclusions

Collapse is one of the main causes of damage to the impervious element of dams.
Collapse is very important in uncompacted materials, specially when non sluiced,
but may also be potentially damaging in high dams with compacted rockfill, spe-
cially when the abutments are irregular and abrupt.
Collapse can be modelled and calculated through the finite element method with
sufficient precision to see the prospects of producing damage in the impervious
element.
In rockfill dams more than 100 m high, rockfill should be compacted in layers not
thicker than 80 em, and in general with vibratory rollers.
141

References

Alberro, J., Leon, J. L. and Guzman, M.A, 1976. "Jose Marfa Morelos". 'Com-
portamiento de Presas Construidas en Mexico'. Comisi6n Federal de Electrici-
dad, Mexico, 127-166.
Baumann, P., 1960. "Cogswell and San Gabriel dams". Trans. ASCE, 125:2A:29-
57.
Bertacchi, P. and Bellotti, R., 1970. "Experimental research on materials for
rockfill dams". 10th World Conf. Large Dams, Montreal, 1:511-529.
Cooke, J.B., 1984. "Progress in rockfill dams". J. Geotech. Eng., ASCE, 110:10:
1381-1414.
Donaghe, R.T. and Cohen, M.W., 1978. "Strength and deformation properties of
rockfill". U.S.Army Engineer Waterways Experiment Station, Technical Report
5-78-1.
Elsden, O. and Keefe, H.G., 19. "Embankment dams-Detail design and construc-
tion". 'Hidroelectric Engineering Practice'. Chapter IX. 'Embankment dams', pp
406-479.
Fumagalli, E., 1969. "Tests on cohesionless materials for rockfill dams". J. Soil
Mech., ASCE, SM1:313-330.
Gomez Navarro, J.L. y Juan-Aracil, J., 1958. "Presas de Embalse". Publications
of the E.T.S. Ingenieros de Caminos, Madrid.
J. Martin Vinas, 1970. "Contribuci6n a la explicaci6n fisicoqufmica del fen6meno
de asiento en escolleras cuando los materiales se ponen en contacto con agua".
Unpublished report.
Justo, J.L., 1968. "Deformaci6n de las Presas de Escollera". Fundaci6n Juan
March, Madrid.
Justo, J.L. and Saura, J., 1983. "Three-dimensional analysis of lnfiernillo dam
during construction and filling of the reservoir". lnt. J. Numerical Methods in
Geomechanics, Wiley, 7:225-243.
Justo, J.L., 1986. "Auscultaci6n y comportamiento de pedraplenes y presas de
materiales sueltos", Simposio sobre Terraplenes y otros Rellenos, Soc. Esp.
Mec. Suelo, Madrid, 315-355.
Justo, J.L., 1988. 'The failure of the impervious facing of Martin Gonzalo sockfill
dam'. 16th Congo Large Dams, San Francisco, Vol. 5, 252-262.
Justo, J.L., Saura, J. and Segovia, F., 1989. "A three-dimensional finite ele-
ment method for the study of the behaviour of embankment dams with thin
earth core during construction and filling of the reservoir". De Mello's Volume,
217-224.
Justo, J.L., 1990. "In situ tests". 'Advances in Rockfill Structures'. NATO ASI,
Lisbon.
Marsal, R.J., Moreno, E., Arenas, A, Guzman, M.A and Saldana, F., 1976. 'La
142

Angostura', in Behaviour of Dams Built in Mexico, Secretarfa de Recursos


Hidraulicos, Mexico, pp. 313-39l.
Naylor, D.J., Tong, S-L and Shahkarami, A.A., 1989. 'Numerical modelling of
saturation shrinkage', Numog III, Niagara Falls (Canada).
Nobari, E.S. and Duncan, J.M., 1972. "Effect of reservoir filling on stresses and
movements in earth and rockfill dams". U.S.Army Engineers Waterways Experi-
ment Station.
Parkin, A.K., 1990. 'Creep of rockfill', Chapter 9 of this book.
Saura, J., 1979. "Estudio tridimensional de tensiones y deformaciones en presas
de materiales sueltos". Ph.D. Thesis. Poly technical University of Madrid, Spain.
Sowers, G.F., Williams, R.e. and Wallace, T.S., 1965. "Compressibility of broken
rock and the settlement of rock fills". Proc. 6th ICSMFE, 2:561-565.
Soydemir, e. and Kjaernsli, B., 1975. "A treatise on the performance of rockfill
dams with unyielding foundations in relation to the design of Storvass dam".
NGI.
Steele, I.e. and Cooke, J.B., 1960. "Rockfill dams: Salt Springs and Lower Bear
River concrete face dams". Trans. ASCE, 125:2A:74-116.
Terzaghi, K., 1960. Discussions, Trans. ASCE, 125,2,139-148.
Valstad, T. and Strcj:m, E., 1975. "Investigations of the mechanical properties of
rockfill for the Svartevann dam, using triaxial, oedometer and plate bearing
tests". NGI, 110:3-8.
Veiga Pinto, A.A., 1983. "Previsao do comportamento estructural de barragens
de enrocamento". LNEC, Tese, Geotecnia.
143

Appendix 1: Parameters of conglomerate in Yeguas Dam, assuming oedometric


conditions (fig. 25 and 26).

The settlements of cells 10 and 16 during construction (fig. 25) were:

slO = 0.601 m

s16 = 0.449 m

And during the rise of water from level 187 to 213.34:

slO = 0.34 m

s16 = 0.078 m

Setlement cells 10 and 16 are aproximately symmetrical respect to the axis of the
dam.
The settlement of cell 16 (downstream) has been used to find a rough estimate/of
the oedometric modulus of conglomerate with the compaction water content (Eoed
= 47 MPa).
The settlement of cell 10 (upstream) with this oedometric modulus would be 0.463
m. The difference between the actual settlement (0.601 m) and the calculated sett-
lement (0.463 m) may me ascribed to collapse and buoyancy.
Collapse and buoyancy during construction:

scb = 0.601 m - 0.463 m = 0.138 m

Equations 16 and 12 may be used to find the oedometric modulus of the satura-
ted material.
From plate loading tests, the unloading modulus of the saturated material was:

ESU = 98.9 MPa

Marsal et al. (1976) have found an average Poisson's ratio of 0.26 (from 0.21 to
0.30) in measurements in sand and gravel during filling of the reservoir.
The oedometric modulus during unloading will be:

1 - \) 1 - 0.26
E~~d = ----------- ESU = ------------------- x 98.9
1- v-2 \)2 1-0.26-2xO.262
144

E~~d = 1.224 x 98.9::: 121 MPa

Equation 16 gives the buoyant heave:

22.7-13.7 1
sb = - ------------ x --- x 13.32 = -0.007 m
121.000 2

Collapse settlement will be:

Sc = scb-sb = 0.138 + 0.007 = 0.145 m

Equation 12 gives:

1 1 13.3
0.145 = (------- - --------)(687 + 22.7x21 + 22.7x-----)x13.3
s
Eoed 47000 2

Eged = 33.8 MPa

Figure 25 shows the lines of equal postconstructive settlement when the reservoir
rose to the level indicated in the figure at the end of June 1988.
The settlement of cell 16 may be ascribed to creep and can be calculated with the
help of equation 23 making HI = 0:

683 + V2 x 22.7 x 30.8


scr = 0.078 = f ---------------------------- x 30.8
47.000

f = 0.115

For estimating the creep settlement of cell 10, we assume a position of the water
level intermediate between the ones of figures 25 and 26:

Vz(187 + 213.34)=200
687 + Vzx22.7x8 687 + 22.7x8 + Vzx13.7x21.4
scr = 0.115(8x-------------------- + 21.4----------------------------------)
47.000 34.000

scr = 0.089 m
145

Buoyancy is given by equation 18:

22.7-137
sb 12 = - ----------- (36.lx30.5 - V2 x 30.5 2 - Vz x 13.32 )
121.000

Sb12 = - 0.041 m

Collapse settlement at cell 10 is:

Sc = 0.34 - 0.089 + 0.041 = 0.292 m


Equation 10 gives:
1 1
0.292 = (------- - ---------)(687 + Vzx22. 7x21 )21
E~ed 47.000

E~ed = 27.5 MPa

Not very different from the one corresponding to the lower layer.
The average Eged is 31 MPa.

Eged = Ege d(1-a)


31 = 47(I-a)

a = 0.35

31
ES = ------- = 25 MPa
1.224
146

Appendix 2: Parameters of rockfill in Martin Gonzalo dam, assuming oedometric


conditions

The set of settlement plates C is placed near the main section of the dam and near
the axis. Figure Ap. 2.1 shows the settlement as a function of log time. Figure Ap.
2.2 shows the log of the rate of settlement as a function of log time.

1100

1000 water level 271


water level 245
900

800
...
c:
CI)

leakage 170 Lis E

...
CI)

E 700 leakage 1000 lis


E CI)
VI

... 600
c= CI)
VI
CI)
a.
E water level 264.2 - - - - - - - - . - - - - - - - - - 1 c
.! 500
...
CI)
o
u

(/) 400
end of construction 8. beginning of impounding

300 facing constructed - - - - - - - - - - : : ; , . . . -

200~ end of fill


100

OL-________~____~____~___L_L!_LI~---L-----L--L-~JI~----L
10 2 4 6 8 10 2 2 4 6 8 103 2
Time ( days)

Figure Ap. 2.1 Settlement versus log time at settlement plates C, level 282.143 m
(Martin Gonzalo Dam).

When impounding starts the rate of settlement increases probably due to the co-
llapse produced in the lower drain by the described leakage.
Leakage increases strongly when the rupture in the membrane is produced. We
147

have shown in figure Ap. 2.1 our estimate of collapse settlement, and in figure Ap.
2.2 the time during which we estimate that collapse settlement has occurred. The
rest is creep settlement.

0> I/)
......
0
0 "C1>
C\I
lOC\I
a.
+-
'" <i- <tl'-

-
ll) +-
L..
0
::l C tD 0 C\IC\I
.....
L..
+- C\I 0
-u'" '"0
C '"
0> C1>
lI)
> >
Cl>

-"
> C1> C1> lI)
0 u
-
Ol

... ...
"0 lI)

--
C 0
0- :::J -'"
0 c L.. 0 CUCl>
0

-
u a. ~ Cl>
0 0
10 c 0 E 0
3: 3: 3:
C1>
8
6 I
4
>-
0
""- 2
E
E
...c
C1>
E 8
C1>
6
:::C1>
I/)
4 -
'0
~
0
a:: 2 -

-I
10
8
6
10 2 4 6 8 102 2 4 6 8 Id 2
Days

Figure Ap. 2.2. Log of rate of settlement as a function of log time at settlement pla-
tes C, level 282.143 m (Martin Gonzalo Dam).

Figure Ap. 2.3 shows the settlement at the settlement plates C as a function of the
height of the dam and time.
At the end of rockfill placement, the maximum settlement, a little above the cen-
tre of the fill, is 1,000 mrn.
148

286 SETTLEMENT
(mm)
000
o 000
r0 {\J

[[]
280
\
\
\
\
\

[]]
270 \
\
\
\
\
\
\

[I]
260 Middle of rockfill E-
o
E .0
o
o
0.
.,
~
I @]~ 10
(.!) o 1-:;;
<.J 1.0
W
I 10.
o I~
., 1~
250 ~ 1u
@J Plate number. I
I
[1] CD End of till placing. I
End of construction and I
\
[] beginning of impounding. I
I
@ Max. water level be fore rupture. I
240
[1] @) Lost readings before rupture I
I
Assumed end of collapse cree.

Latest reddings.
[IJ Rock 0 Theoretical parabolic settle-
::::;::0'" ........... 7'" ----:: ment at end of till placing.
230
500 1000 1500
SETTLEMENT (mm)

Figure Ap. 2.3. Martin Gonzalo Dam. Settlement versus height at different dates.
Set of plates C.
149

Using a one-dimensional interpretation of settlements (a three-dimensional calcu-


lation is mentioned in chapter 7), the oedometric modulus during construction
may be obtained from the equation:

yH2
Eged = --------- CAp. 2.1)
4 smax

d 22.5 x 53.03 2
Eoed = ----------------- = 15,788 kPa-::::-15.8 MPa
4x1

The theoretical parabolic settlement along the height of rockfill is given by the
equation:

y (H-z)z
s = ----------- (Ap.2.2)
z d
Eoed

We may see in figure Ap. 2.3 that the agreement between measured and calcu-
lated settlements at the end of fill placement is fair.
From the end of fill placement till the beginning of impounding there is creep,
represented to the right of figure Ap. 2.3.
We have also represented there the theoretical parabolic creep according to equa-
tion 25.
Up to a little more than half the height of the dam, creep is better represented by
a straight line. This means that the vertical creep of a layer is a percentage of the
height of the layer, independent of the pressure at which the layer is subject. This
"creep settlement" is, in figure 31, a 0.28% .
In the upper half, the creep settlement has the parabolic shape corresponding to
equation 25. If we apply equation 24 to the upper half of the dam, we have:

fx22.5
0.163 = -------- x 14.12 x 20.52
15,800

The creep coefficient is:

f = 0040

We represent in figure Ap. 2.4 the settlement occurring between the 2nd of Fe-
150

bruary 1987 (beginning of impounding) and the 6th April 1988, when the sett-
lement rate decreases to the level existing before leakage.

Crest of dam
286r-------------------__

@
280

o
272 (wate level d urino year 90)
270

E
260 the dam

250 5

240 Plate number.

Settlement from 2-ll- 87 till 6-I\[-88


+ Total post- ti II settlement
Fundation 0 Creepaccording to eq 25
7"'-/-

230L-~--------~----------~---
100 500 1000
Settlement (mm)

Fig. Ap. 2.4 Martin Gonzalo Dam. Creep and collapse settlement versus height. Set
of plates C.
151

As indicated in figure 28 the level reached by water leaking at the axis of the dam
was 240 m. We see that although the rate of settlement at plate 11 strongly increa-
sed during this period, "most of the settlement during the so-called collapse pe-
riod" occurred above the water level.
Considering only the settlement below level 240, equations 12 and 16 may be used
to evaluate collapse.
The average pressure at the centre of the layer suffering collapse is:
a av = 49.52 x 22.5 = 1114 kPa before water rise

a 'av = 46 x 22.5 + 3.52 x 13.0 = 1081 kPa after water rise

Assuming 'J= 0.3:


su
Eoed = 235 x 1.346 = 316 MPa

Bouyant heave (eq. 16):

22.5-12.0
sb = - ------------ X V2 x 7.03 2 = -0.001 m negligible
316,000

Collapse settlement (ep. 12):

1 1
0.0086 = (------- - ---------) x 1114 x 7.03
E~ed 15,800

E~ed = 13.5 MPa

13.5 = 15.8 (I-a)

a = 0.146

We see that collapse is quite moderate.


We have represented also in the same figure the whole post-fill settlement. It may
be composed of two parts:
1. Up to half the height of the rockfill a straight line with a slope of 0.96% the
height of the dam.
2. In the upper half a parabola corresponding to a "creep coeficient" f given by
equation 24:
152

fx22.5
0.887 - 0.255 = --------- x 15.18 x 22.64
15.800

f = 1.29

We have included also, in figure Ap. 2.4, creep according to equation 25. We see
that the agreement with measured creep is fair.
The influence of water loading, not evaluated here, seems to be quite small.
CHAPTER 7
TEST FILLS AND IN SITU TESTS
1. 1... JUSTO

1. Why in situ tests?

In situ test is the best established method fo r compaction cont rol in any kind of soil.
Laboratory compaction tests cannot usually be carried out in rockfill , because the
large sizes of rock cannot be introduced in the laboratory moulds. Usually it is not
allowable to employ sizes larger than one fifth the smaller inner dimension of the
container (v. Salas and Justo, 1975).
The use of s maller sizes and correctio ns is commonly not a llowable when the
~coarse" particles (nol introduced in the mould) are more than about 30% of the
total weight of roeHiIl, because the "fines" do not compact well insid e the skel e-
ton of the coarse particles in the field.
Thi s means that common laboratory compaction moulds can never be used to
compact true rockfi ll.
For that reason, in rockfiJI , in situ te sts must be used not only as a compaction
control, but also as a compaction reference.
There are three fill properties relevant in civil engineering works: permeability,
shear strength and deformability. The last two have a direct type relat ionship with
density that is usually, although not always in rockfi ll, more easy to measure.
It is interesting to remember the opinion of Sherad et al. (1963) respect to field
tests:
~Because of the large size of the individual rock particles, it lias not been con.side
red practicable or necessary on. most jobs to carry out any testing f or comrol during
cOltstrnctioff. At a few dams large-scale field density tests have been made by exca-
vating a large hole in the fill and measuring the volume of the hole by surveying.
For the ex te nsive tests at Goschenenalp dam, it was found necessary to dig out
about 190 m3 of rock in o rder to assure that the density cou ld be measured with an
accurayof 1%. The porosity (or densi ty) can be roughly estimated and often is, alt-
hough the value obtai ned is of little be nefit to the engineer and would nOl be of

'"
E. MoronilO dol NtlYS (M.). Adl'Onn~.s in RIXk/i1l Strucmrn. 153--193.
C 1991 KI_v:r Acodt mic Ptthli.shtrs.
154

much additional help even ifit were and exact quantily, the estimate is made by mea-
suring (1) the quantity of rock used (either by weighing it in trucks as it is delivered to
the dam or by measuring the volume in the quarry) and (2) the volume of the rockfill
section.
The porosity generally varies between 18 and 35%. The numerical value depends
primarily on the gradation of the rock and the percentage of the fines and only se-
condarily on the degree of compactness of the fill. In general, the higher the con-
tent of fines, the higher the density of the embankment and the lower the porosity.
The most valuable field testing on rockfill sections is that which is carried out to
determine the relative influence of the various types of compaction equipment and
effort. Since measurements of the density are almost impossible to make, the re-
sults of tests of various types of compactors can most easily be measured by leve-
lling the surface of the layer.
A method that has proved practical is to paint the exposed points of a number of
rocks on the surface of the layer being studied and to measure their elevations be-
fore and after compaction. The average difference in elevation gives and indicati-
on of the change in the density. With such field tests it is possible to study the in-
fluence of the thickness of the layer, the type and number of roller passes, and of
sluicing of the fill. For a given roller and layer thickness, we can determine the
number of passes at which additional coverages no longer give any large increase in
the density. However, these tests still do not give any assistance in evaluating the
problem of determining the influence of the method of compaction on the com-
pressibility of the embankment".
Up to here the quotation from Sherard et al.
We shall review the different kinds of in situ tests, and comment on them. Due to
the limitation of laboratory tests, the reference for in situ tests is usually obtained
from test fills.

2. Test fills

Fig. 1 shows the layout of one of the two test fills formed before the construction of
Martin Gonzalo rockfill dam. The plan dimension is 20x30 m (slopes not include-
d). According to Bertram (1973) the minimum size of a rectangular test fill should
be 9x18 m.
As seen in the figure, the test fill is longitudinally divided in two halves of lOx30
cm. In each half a different type of rock was employed.
In each half there are 35 measuring points (Bertram recommends around 25),
allowing to draw five transverse and one longitudinal profiles.
155

11
I I
---, f -
12~3f45 - T-'
3.00 345

i2 _.L-._ - - T-3
~
2
3.00
~~~EfJ-3~~
[3D
~

3.00
-,
3.00
14 4
~
~
-f2l-~ts-~ ~-~~.- - T-5
30.00 3.00

--e-r-
i6 !6
~

3.00
~
~t~~ ~ T-7

_t_
3.00
-. 8
3.00
12~9 4-;- '- T-9
1
~
~

12 345
3.00
- ~
!IO 110

~I I(f)

FIRST LAYER (LAYOUT OF IN SITU TESTSl.

o LOADING TEST.

ca PERMEABILITY TEST.

o DENSITY TEST.

Cl WETTING MEASUREMET.

Figure 1. Layout of test fill with blown rockfill (Martin Gonzalo rockfill dam).

Side slopes of 1 vertical on 1.5 horizontal (fig. 2) are recommended, with ramps at
each end having 1 on 5 slopes for convenient travel of the tractor and roller.
Six layers were provided in each fill with the following thicknesses:
- 2 layers of 40 cm
- 2 layers of 60 cm
- 2 layers of 80 cm
Water was sprinkled from tank trucks after the layer had been extended (fig. 3) up
to the moment when water flowed through the base of the fill. That happened when
the volume of the water added was around 10.7% of the volume of the fill.
A time was waited for air-drying of the material before compaction.
Compaction was carried out with a Dynapac vibratory roller(fig. 4), with a mass of
12 Mg.
156

Figure 2. Martin Gonzalo. General view of test fill of ripped material.

Levellings were taken each two passes of the roller.


According to Bertram (1973) to obtain a representative reading at each grid point
it is advisable to set the level rod on a 30 cm square steel plate having a raised stud
in the center on which the rod can be held (fig. 5).
Figure 6 shows the longitudinal and transverse profiles obtained by sucessive leve-
llings (fig. 7), corresponding to the test fill of slate, third layer.
A thorough study of the compaction was carried out, and the following conclu-
sions were reached:
1. When the layers thickness decreased from 80 to 40 cm there was no clear incr-
ease in percent compaction.
2. When the number of passes increased, the compaction increased, but this is the
result of particle breakage, and might impair the permeability of the rockfill.
In the test fill of Mont Cenis Dam, the number of passes ranged from 2 to 16, and
although the settlement continued to increased for 16 passes, it was checked that
the settlement produced by the first four passes was double the settlement produ-
ced by the following four. It was concluded that the technical-economical optimum
stayed between 6 and 8 passes.
157

Figure 3. Watering of graywacke rockfill with tank trucks.

In Goschenenalp Dam test fill, dumped rockfill (actually a morraine) reached a


dry density of 2,250 kg/m3 (porosity 17 to 18%). Com~action with a vibrating ro-
ller of 15 Mg only increased the density to 2,300 kg/m . The grain size curve was
very well graded. As a result the shells of this central earth core dam, 155 m high,
were not compacted, but dumped in 2.5 m lifts.
Grain size tests of the compacted material were performed, to study particle
breakage, with the results indicated in figure 8. Settlement plates were also placed
to study the settlement of layers at depth (fig. 7, 9 and 10). The problem with these
plates is to maintain them free from damage by the compaction equipment.
Examination of the appearance of each layer of a test fill may be important for a
proper evaluation of the effect of the compaction. After the test fill has been com-
pleted, an excavation may be made to study, for example, segregation of fines.
Figure 11 shows the trench made at New Melones dam. At Martin Gonzalo test fill,
excavations 1x1 m in plan, and with the depth of the layer were made to check the
penetration of water through the layer. The penetration was correct.
In Cougar sloping core rockfill dam, comparative tests showed the advantage of a
9.1 Mg vibratory roller over a 4.5 Mg vibratory roller, and of this over a 45.4 Mg
rubber-tired roller.
158

Figure 4. Twelve megagrams vibratory roller.

Bertram, after examining the results of test fills at the U.S.A., recommends the
use of vibrating rollers from 8 to 10 Mg, and perhaps larger, and layers from 60 to
90 cm. The same conclusion was reached by other authors.
He indicates also that the vibratory roller is not effective to compact rockfill with
an excess of fines. The frequency of the vibratory roller is another aspect that des-
erves consideration. At Venemo rockfill dam it was found that with 1,000 or 1,800
r.p.m. the settlement was larger than with 1,250.
The following in sity tests were carried out at the Martin Gonzalo test fill:
1. Plate loading tests under compaction moisture conditions or under saturation.
2. In situ density
3. Permeability tests.

3. Plate loading tests

A good rockfill needs usually high shear strength, and it generally fulfills this condi-
tion.
159

Figure 5. Level rod on steel plate (Bertram, 1973).

When employed in a dam shell, at least as important as that is to have a high per-
meability, and this condition is, also, generally accomplished.
Finally when employed in a road or a dam it needs an acceptable deformation
modulus. This is the most critical property for rockfill, because the lack of it might
impair the safety or the economy of the dam, through the need for freeboard, or
might produce the cracking of the impervious element or pavement.
Several laboratory methods are employed to find this modulus, and for field tests,
plate loading test.> have also been used.
As far as we know, the first plate loading tests on rockfill were performed by
Guilhamon and Castelnau (v. Justo 1968) around 1954 to check the effect of diff-
erent types of compactors on the rockfill of Iril Emda Dam, in Algier. The grain
size curve of the rockfill is indicated in figure 12.
The results reached are shown in table I.
160

Table I. Plate loading tests at Iril-Emda dam (Guilhamon and Castelnau, 1954).
Compaction

No compaction 5
Many passes of 21 Mg truck 30-40
1 Mg hand compactor 25
200 kg vibratory roller 25
15 Mg smooth roller High for small loads
10-15 for larger loads
1600 kg vibratory sledge 35 to 80 according to
m.mber of passes

LONGITUDINAL PROFILE
1 2 3 4 5 6 7 8 9 10
0
1
-----'-
..... ~ ..........
-
- .....................
2
"-
.... _-----
3
4
5
E
0 0
I-
Z
w 2
~ 3
w '-'",
4
..J ......
I-
I-
5 '- '-,

w 6
C/)
'-.
7 _ _ _ UNCOMPACTED MATERIAL.
8 - - _ 2 COVERAGES OF COM PACTO R.
9 -._.-4
10 ____ 6
II
" "
"
I t t I
11

Figure 6. Settlements produced by compaction. Slate test fill. Layer thickness 0.6 m

Plate loading tests and settlement measurements at dam show that the modulus of
deformation is very high (200 MPa at least) for pressures less than 400 kPa. When
the pressure increases from 400 to 700 kPa, the modulus of deformation decreases
up to 50 MPa. Finally, when the pressure increases from 700 to 2,000 kPa, the mo-
161

dulus increases a little (60 MPa).

Figure 7. Levelling of slate and watering of graywacke rockfill.

This reveals the existence of an overconsolidation produced by compaction, that


makes difficult to find the modulus of deformation from laboratory tests.
Plate loading tests were also made in Svartevann Dam (Valstad and Stllml, 1976).
The maximum size of rock was about 1 m. The plates were circular with diameters
of 500 and 750 mm. The average E value for 500 mm plates was 88 MPa, and for
750 mm plates 109 MPa, both during first loading. The average value for reloading
(first and second) was 292 MPa.
The values obtained from the triaxial and oedometer tests were in reasonable
agreement with the results from the plate bearing tests.
In Martin Gonzalo test fill more than 85 plate loading tests were carried out with
circular plates with diameters 30, 45 and 60 cm.
The tests were carried out under compaction wetting conditions (fig. 13) or with
the soil submerged (fig. 14 and 15). First an excavation was made to the centre of
the layer.
A detailed description of the tests has been made by Del Campo (1984), Justo et
al. (1985) and Justo et al. (1988).
162

100
I
I
90
I
80
/
70
(!)
60
/ I
~.

z
(f)
k/
/ II
(f)
50
/1/
<t /
a... /
40
0~
30
// J:/ /
/ V/
.........-< ..-
II
20 '"
10
~
V V
1..-/
/~
W
o ~
~
---- .- .-
~

0.06 0.2 0.6 2 6 20 60 200


GRAIN SIZE (mm)

Figure 8. Particle breakage in Martin Gonzalo rockfill.


A. Material unloaded by trucks
B. Compacted material.

Figure 16 shows the pressure-settlement curve for a plate loading test. The rockfill
was saturated after the settlements were stabilized under a pressure of 500 kPa.
From that point on a decrease in the modulus of deformation is observed.
The load was applied with a hydraulic jack (fig. 13 to 15), using as a reaction a
loaded truck, and is mesured with a manometer (fig. 12 and 13).
Settlements are measured with three dial gauges (fig. 14 and 15) whose average is
used to find the pressure settlement curve. The dial gauges are joined to bars who-
se support is far away from the plate, and rest on it.
The plate is placed on a horizontal levelled surface, interposing a fine layer of
normalized sand. Once the plate is levelled, an "initial" presure of 20 kPa is applied
to assure a uniform support, that corresponds to the zero of settlements.
The load was applied in increments of 50 kPa up to 250 kPa, and then in incre-
ments of 100 kPa. Unloading was made with decrements of 500 kPa. A new
increment was applied when the settlement in the previous loading interval is less
than 0.05 mm in 2 minutes.
163

pipes stoppers
--+
150
-/ ~/

\
\~l11Jlankment surface

25 mm galvanized pipe, in 1 m
--'--_._-
--- --- ,---

sections, with threads at both ends

100 mmfl-'-!J.E~Lin 1m section~_with


threads at both ends

Foam rubber ring

"'(.

SEC T ION A - A'

To
I

I
Galvanized steel plate
- r---( 6 mm t hi c k )

W
o_~ __
o A
I
Measures in mm

-+
-+-- ~QQ_--+
PLAN
!

VIEW

Figure 9. Settlement plate.


164

The modulus of deformation is calculated from the formula corresponding to the


elastic settlement of a rigid plate on Boussinesq space:
IT p(kPa)-20
E(MPa) = ---- (14) -------------- D (cm) (1)
4 s(mm)
where:
p = pressure
D = plate diameter
s = settlement

Figure 10. Settlement plates in slate rockfill.

Valstad and Str.0'm (1976) have found a values of Poisson's ratio during oedo-
meter test in rockfill of 0.2, little dependent upon porosity.
A summary of the tests is presented in tables II and III.
There was a wide sacattering in the moduli obtained during loading, which ran-
ged from 11 to 84 MPa, with an average value of 24 MPa for the compaction water
content. The plate loading tests were also continued, for compactation control, du-
ring dam construction.
El corresponds to the "initial" pressure to 500 kPa. The pressure interval of E2 is
165

Table I I. Average values of modulus of linear deformation obtained from plate loading tests
in Martin Gonzalo rockfill dam. In brackets number of tests ( = 0.3)
,------ I I I
I I IFirst loadingl Reloading
I lIater INo. ofl Plate I
ILocation Rockfi III 1 I 1 E1 E2 E E4 Layer
3
1 1 content 1tests Idi ameter 1
1 I 1 MPa MPa MPa MPa
1 1 1
1 1 1
1Test fi II ripped 1compact i on 1 2 30 cm 27
1 1 1
ITest fi II ripped Isaturated 1 3 30 cm 19
1 1 1
ITest fi II blown Icompact i on 1 12 30 cm 22 30
1 1 1
Test fi II blown Isaturated 1 12 30 cm 22
I 1
Dam blown Icompactionl 10 30 cm 20 32 59* 275* 1st to 4th
I 1
Dam blown Isaturated I 30 cm 13 22 51* 235* 1st to 4th
1 I
Dam blown Icompact i on I 14 45 cm 33 84 5th to 9th
1 I
Dam blown Isaturated I 2 45 cm 55 80 6th
I I
Dam blown Icompact i on I 9 45 cm 81 127 6th to 9th
1 I
Dam blown Icompactionl 60 cm 116 (8) 11 (7) 12th to 24thl
1 1 1 1
Dam blown Icompactionl 6 60 cm 1 56 35 12th to 24thl
I 1 I 1
Dam blown Icompaction 1 40 cm 123( 13) 15(11) 87 (10) 49 (9)123rd to 27thl
I I square 1 I 1
I I 1 I I
1compact i on I 123(59) 124(54) I 60*(10)1 275*(10)1 I
IAverages I I I 1 1 I 1 I
I 1compact i on I I 1 1 75 (25)1 56 (24)1 I
I 1 1 I 1 1 I I I
I 1saturated 1 115 (8)1 22 (17)1 51*(5) 1235*(5) I I
I 1 I 1 1 I 55 (2) I 80 (2) I I

*Unloading
166

from 500 to 1000 kPa.


The following conclusions were drawn from table III:
1. The modulus of deformation during reloading was from 2.5 to 4 times larger
than during first loading.
2. The modulus for unloading is very large in the first part of the unloading inte-
rval (1000 to 500 kPa).
A 12 Mg vibratory roller was used for compaction of the test fill. No important
change in the moduli was found when the thickness of the layers increased from 40
to 80 cm, or the number of passes from 6 to 8.

Table III. Summary of moduli of deformation of rockfill from plate loading tests in
fill test and dam.
Loading Reloading Unloading
Water
Rockfill E11 E21 E1r E2r E1u E2u
content
(MPa) (MPa) (MPa) (MPa) (MPa) (Mpa)

as compacted 28
Ripped
saturated 19

Blasted as compacted 24 75 56 59 275


&
ripped satured 15 I 22 55 80 51 235

In the dam it was decided to compact with 6 passes of a 14 Mg vibrating roller in


80 em-thick layers. The rockfill was duly watered during construccion.
Settlement plates were placed at the dam during construction. To obtain a good
agreement between measured and calculated settlements, a modulus of deforma-
tion of 13.2 MPa in the 11 lower meters of dam, and 9 MPa in the upper 44 m had
to be assumed (fig. 19). The average modulus to fit the measured settlements is
about 40 % the average modulus from plate loading tests. No clear explanation for
this can be given; it must be stated that the duration of the test in plate loading test
(somewhat more than two hours) does not allow creep to develop. Segovia (1989)
has evaluated the importance of creep in the deformation of dams.
In the conglomerate (fig. 18) of Yeguas dam, whose grain size curve is indicated
in figure 19, the modulus from plate loading tests (45 em diameter) was somewhat
larger than the one measured in the dam during construction.
167

Figure 11. Test fill at New Melones Dam, excavated for inspection (Bertram, 1973).

4. In situ density

We have already quoted in section 1 Sherard's comments on density measuremen-


ts.
In Martin Gonzalo, in situ density by the sand method was carried out (fig. 20).
This is only possible when the size of rock is small.
Bertram (1973) indicates that field density tests on rockfill are cumbersome and
time consuming because the size of both the sample and the equipment is very lar-
ge.
Figure 21 shows a method to measure dry density in the field, in rockfill with large
sizes of rock:
1. First the surface of rockfill is carefully levelled.
2. A big hole, relatively flat and wide, with a regular botton, is excavated. The ma-
terial is carefully placed in a container to be weighted. The water content, w, is
measured, and the dry mass is found through the formula:
168

M
Md = ------- (2)
l+w

I ____/ /
100r--------------.--------------~------~------~

90r-------------~--------------~~~--~~
/------~
/ //
80~-------------r----------~~~-,~--------~
/ //
~ 70r---------------4----------;I~~--~/v~---------------'
~ 60r----------------~-------;I~------4_--------------~
~ 50~------------~----J7~~----/~/~------------~
~ 40~------------4--/~L-----/~/~~------------~
30r--------------4~~------1----+--------------~
/V /1
20r------- ~~L4------~---+----------~

----
_____ /1
10r---~~~----~----~~-------+--------------~
~-----
O~~~--------~---------------L--------------~
I Ix 10 Ixlo2 Ixl03
PARTICLE SIZE (mm)

Figure 12. Grain size limits for Iril Emda rockfill.

3. The hole is carefully covered with a plastic sheet. The plastic is filled with water
up to the level of the ground to find the volume of the hole V.
4. The dry density will be:
Md
P d = ----- (3)
V
Density measurements were also carried out in connection with permeability tests
inside a tube, as described in next section.
In Martin Gonzalo test fill with ripped material the densities reached by nuclear
methods, sand method and tube method were compared (v. Justo and Manzanares,
1983). The largest density was reached by nuclear methods; the sand method gave a
density of 92% of the former, and the tube method around 91 % the density of the
sand method.
169

Figure 13. Plate loading test under compaction wetting condicion.

Figure 22 shows the relationship between the oedometric modulus, in ordinates,


and dry density in abcissae obtained from several laboratory tests on granular
materials.
We see first that for a given dry density the oedometric modulus may range from
20 to 110 MPa.
The points which are joined by a line correspond to the same material with diffe-
rent compaction degrees. We see that an increase in density produces often a large
increase in oedometric modulus.
In this figure the materials with rough surface, represented by small circle (dry) or
latin crosses (wet material) correspond to rockfill.
As a rule, the rounder the particles and the more extended the grain size curve,
the larger the density for a given compaction energy, but this does not mean that
the oedometric modulus will be larger.
Table IV shows the relationship between oedometric modulus deduced from sett-
lement plate measurements during construction and dry density at some rockfill
dams made of presumably soft rockfill. We see that there is no relationship bet-
ween both quantities.
The same is shown in table V for concrete face rockfill dams.
170

Figure 14. Plate loading test under submerged condition.

5. Permeability test

As indicated above, when rockfill is to be used as a dam shell it is very important


that it be free draining, because in this case the stability of the dam during cons-
truction, during rapid drawdown or with high water level is greatly increased.
In the case of cracks appearing in the impervious element the dam is also much
safer if the downstream rockfill fulfills that condition.
This tests is only needed in poor rockfill.
In Martin Gonzalo permeability tests were carried out in the following way:
Once a layer is compacted, an excavation lxlxO.25 m is made and filled with wa-
ter.
In some cases it was not possible to fill the excavation, due to the high permea-
bility of the rockfill, although the discharge of the hose was 2000 liters per hour.
Once the excavation was full, the descent of the water level was measured as a
function of time (fig. 23).
171

Figure 15. Plate loading test under submerged condition.

One method for an approximate determination of the coefficient of permeability


is indicated in Appendix 1.
With ripped rockfill another technique was employed: a tube with an inner diame-
ter of 20.5 cm was sunk about 40 cm in the test fill, and the inside of the tube was
cleaned of rock (fig. 24). A constant discharge test was carried out. According to
the Earth Manual (Bureau of Reclamation), the coefficient of permeability is given
by the relationship:

Q
k = --------- (4)
5.5ro H

where:
Q = discharge
ro = inner tube diameter
H = height of water inside the tube
172

--
11

10

.... l.----'"
~
- V
~
::;.-

/
9 ~

(.!) /
-Clz
/
E 7
E
0
I- 6 0
z ...J

/
w u..

v
~
w 5
...J
l-
I-
W 4
/
(/)
V
3
/
/
2 /
/
/

./
V
a 100 200 300 400 5 00 600 700 800 900 1000
PRESSURE (kPa)

Figure 16. Plate loading test. Pressure-settlement curve.

The coefficient of permeability can better be obtained from a falling head per-
meability test, according to the formula (v. Schmid, 1966):
ro In hd h 2
k = ---- -------------------------------- (5)
4 3(hrh2)
t2[------------ + 1]1/3 - t1
4t,srnr o

where:
173

r 0 = inner radio of tube


h 1 = initial height of water in tube
h2 = final height of water in tube
lSr = increment in the degree of saturation produced by wetting
t 1 = initial time, corresponding to a wetting front at a distance equal to r 0
from the base centre.

@v CD @
55 - 55 ~

50 -
E
E SO
t- 40 - ~ 40
:::c S2 " ... ... , ,
t!>
30 W 30 - \

W :::c
:::c .-l /
/
/
~ 20
/
.-l
20 u.. "
.-l
G: ::.:::
'"
... ,.?'
::.::: 10 E =13.2 Mpa (h<l1 m) ulO ~

u 0
0
E = 9.0 Mpa (h>ll m) 0:: @
0:: , vC5) I I

0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0 1.2

S ETTL EMENT (m) SETTLEMENT ( m)

Figure 17. Comparison of measured and calculated settlements at Martfn Gonzalo


rockfill dam.
1. Evolution of settlements with rockfill height
2. Total settlements at the end of construction
A = measured settlements at plates
B = calculated settlements at plates
C = top of dam
D = base of dam.
174

Table IV. Oedometric moduli measured, during construction in dams made of pre-
sumably soft rockfill.

Pd qu Eoed
Dam Rock
3
(Kgjm ) (MPa) (MPa)

Serpentine quartzite & schist 110-150

Eggberg friable sandstone & 75


weathered gneiss

Mackintosh slabby slate & 63


graywacke

Bigge shale & clayey 44


graywacke
Scarrrnonden friable sandstone 2020 55-80 42

Steinbach weathered schist 36

Scotts Peak very soft argillite 2100 8-30 34

Llyn Brianne slaty mudstone 3250 35-100 29-33

Balderhead shale 1840 29

~l nscar carboniferous 2030 25


sandstone

Martin Gonzalo slate & graywacke 2140 9.8

Muddy Run sound to weathered 2000 9.8


micaceous chist

Weighing the material taken from inside the tube and dividing by the volume co-
rresponding to the average diameter of the tube, a density test was carried out.
According to Schmid (1966), equation 5 gives a lower bound of the permeability.
The results reached by application of equations 4 and 5 are compared in Appen-
dix 2.
175

Table V. Oedometric moduli measured, during construction, in concrete face rock-


fill dams.

Eoed n
Dam Pd Placement
(MPa) % ( kg/m3 )

Murchinson 220 11.5 2300 c


Bastyan 156 c
Lower Pieman 156 c
Alto Anchicaya 100-170 23 2280 c
( 130)
Serpentine 110-150 c
Cethana 110-180 2070 c
( 126)
I r il-Emda 50-200 20.3 2150 c
(115)
Wi lmot 96 c
Tullabardine 94 c
Paloona 70-90 c
Quoich 68 27.1 c
Mackintosh 63 c
Rama 44-82 2100 c
Candes 58 c
Naussac 57 2080 c
Foz de Aeia 33 25 2120 c
Contrada Sabetta 28 d

Bou Hanifia 17 26_5 2280 h

c = compacted
h derrick or hand placed
d = dumped
176

Figure 18. Conglomerate of Yeguas Dam.

6. Tension tests (Uriel and Perez, 1981).

In Canales rockfill dam there was an abrupt change in one abutment (fig. 25).
Finite element computations showed the existence of high tension strains in the
upper part of the clay core (fig. 25 and 26), above that position and in the neigh-
bourhood of a steep abutment.
A test fill (fig. 27) was constructed with core material above a flat jack placed
across the base. This jack is composed of three flexible tubes, 20 cm in diameter
when inflated by water pressure. The three tubes are protected inside PVC sheets
and covered by mild steel plates, 3 mm thick. These plates distribute the jack pres-
sure upwards to the compacted clay embankment fill, until the tensile stresses in-
duced in the surface produce the cracking in the fill (fig. 28).
The horizontal strains are measured between fixed points (fig. 28) by means of
pirex rods and dial gauges with 0.01 mm precision. The vertical deflections are
measured by means of two rows of levelling marks placed at the same reference
points (fig. 27 and 28), 50 cm apart in the central part and 75 cm in the outer zo-
ne.
177

100

90

80

70

(.!) 60
z
(f)
(f)
<t 50
a..

0~
40

30

20

10

0
100 10 0.1
PARTICLE SIZ E ( mm)

Figure 19. Grain size curve of the conglomerate of Yeguas Dam.

In order to avoid moisture content losses during the test, the clay is covered with
parafine oil.
Figure 29 shows the vertical displacement of the reference points for different
jack pressures.
In figure 30 the strains, either tensile or compressive, between each pair of mea-
suring points, are plotted against jack pressure.
The first crack appeared between points 5 and 6 for a jack pressure of 340 kPa. At
this moment the unit tensile strain between such points was 0.24%.
The measurements were analyzed by the finite element method, and the follo-
wing conclusions were reached:
1. Up to a tensile strain of 5 x 10- 4 the stress-strain behaviour was linear.
The modulus of deformation in the tension zone was in the range 24-38 MPa, and
for the compression zone 13-30 MPa.
2. When the first crack appeared the mean modulus was 12 MPa
3. The field cracking strain (0.24%) was inside the range of the laboratory test re-
178

sults, although in this case the scattering was large.

Figure 20. In situ density test by the sand method in the top suface of rockfill.

The duration of the standard test was 9 hours.


As the extension strains calculated in figure 26 are larger than the cracking strain,
it was decided to reinforce the tensile zones of the core with a geomesh in the left-
hand side of figure 24, and to place a preload on the right-hand side.

7. Shear strength tests

It is not easy to carry out in situ shear strength tests in rockfill.


Figure 31 shows a plate loading test carried up to failure. The reaction necessary
is about 80 Mg.
The first problem in cases in which a maximum presure is nost reached is how to
define failure (v. De beer, 1970).
179

Figure 21. In situ density test by the plastic method. (Torres Novas-Fatima rockfill
embankment, Portugal).

Figure 31 shows the bearing pressure, qf, according to the authors.


As we know, the bearing pressure is, in the case of a plate on the surface of the
ground:

(6)

For a circular or square plate s = 0.6. Substituting in equation 6:


3.3 qf
N.yC cp') = --------- (7)
yD

The second problem is the evaluation of Ny as a function of cp', because there are
several expressions (v. De Beer, 1970). The following expression is recommended
(v. Jimenez Salas & Justo, 1981; Brinch Hausen, 1970):
tan'" 2 1T ,
N y = 1.5 [e1T 'I' tan (--- + ---)-1] tan ' (8)
4 2
180

120
Eoed= 168 M Po
~ d= 1760 kg/~
110

100 11
90

c 80
a..
~

en 70
::::>
-.J
::::>
0 60
0
~

C,;)
50
a::
t-
w
::!:
0 40
0
w

30

20

10

0
0 o o o o o o
o(\J o o o o o
0
0 v CD eX)
o(\J (\J
(\J

DRY DENSITY ( kIJ/m 3)


Figure 22. Relationship between oedometric modulus (for a pressure increment
from 0 to 2350 kPa) and dry density, according to several laboratory tests on
granular materials (Salas and Justo, 1971).
The ciphers indicated in the graphic are uniformity coefficients.
Rounded and dry particles )( Angular and dry particles 0
Semiangular and dry particles 0 Angular and wet particles +
The same material dry in one case and wet in other o~
181

Figure 23. Infiltration test (Martin Gonzalo test fill).

This expression allows to find </>'.


The average value of </>' found in this way for Svartevann dam was 53, for an ave-
rage bearing pressure of 3700 kPa and a porosity of 26%. This test gives an angle
much higher than the triaxial test.
One inconvenient of this test, as a shear test, is that the stress level is too high for
studies of slope stability in rockfill dams. The advantage is that the test may provide
both: the modulus of deformation and the angle of internal friction.
A passive failure test has been devised by S. Uriel and applied to field measu-
rements in at least four Spanish rockfill dams, among them La Vifmela Dam. A
sketch of the test is presented in figure 32.
In situ direct shear tests on rockfill or boulders with a maximum size of 200 mm
are described by Jain and Gupta (1974) and Ranjan and Prakash (1975).
The set-up, although ingenious, is rather complicated, and hardly justifiable except
for very special cases. In any case the maximum size of rock that may be used is
insufficient for most rockfills.
In a crushed shale the results reached were:

c' = 32 - 80 kPa
182

' = 28.6

Figure 24. Tubes driven in ripped slate rockfill for permeability tests.

8. Conclusions

Owing to the large size of rockfill, in situ tests are invaluable tools to find the pa-
rameters for calculations and check the compaction.
Although both, laboratory and in situ tests have their place in rockfill testing, we
183

believe that there is a trend for an increasing use of the latter, and that this trend
will be reinforced in the future.

Figure 25. Longitudinal profile of Canales dam, showing zones of possible tensile
cracking (Bravo et al.,1988)
a & b tensile strain zones.
o 0 0 0

2t-
~
~ ~
If) If)
fl.
If)

~lllt.II.'
o 0 0 0000

~ ~~~~~~
If) If) II) II) II) II) II)
N r<i -i.ol!i1'-CO
. .
1\ I\, ~ 0:.TI
~\..v I
" 1--. .......
H
"017
CS'I
77
~
""'I"'-/V~ ~/\/~I v"v'\/V\;'\/\/'r17
~/
V
""'/
Vl
Figure 26. Isolines of tensile strain for a crest settlement of 1.5 m. For situation of a
& b see figure 25 (Bravo et al.,1988).
184

11. 5 m
1 1

,I I, ,
I I,

I I I
1 2 3 4' 5 7 8 9 10
ROW I _ _......H'He;'-.-~.-..
e
" , I 5m 8.5m
,I,
ROW II ........t-ff--ei--4. . .~........-e
TWO SUPERFICIAL I I ,
II I
EXTENSOMETERS ROWS I I

THREE FLEXIBLE
PLAN VIEW
MATERIAL TUBES

I.. 8m

1 2345678910

PRECISSION TOPOGRAPHIC
LEVELLING DEVICES 2.5m
STEEL PLATE

C ROSS-SECTIO N THREE FLEXIBLE


MATERIAL TUBES

Figure 27. Test fill for tension test of core at Canales dam (Uriel and Perez, 1981).

For both types of test, more reserarch is needed to find the true relationship bet-
ween the parameters found in the tests and the ones that would be relevant for the
calculation of the dam. This should be obtained from more comparisons between
calculated and measured deflections (and perhaps stresses) in the dam.
185

Figure 28. Cracking in test fill for tension test of core material (in rockfill dam).
Observe strain meters and levelling marks (courtesy of M. Alonso Franco).

E 20
~
f-
Z
w 15
:::!'
w
U
<{
--'
0- 10
~
0

;i 5
u
f-
a:
w
> 0

JACK PRESSURE JACK PRESSURE


1 kg/em 2.50 kg/em - - -
1.15 2.70
1.50 3
1.75 3.25
2 3.45
2.10 3.50

Figure 29. Vertical displacements at reference points (Uriel and Perez, 1981).
186

0
4
I I
REFERENCE POINTS NUMBER
I 2 81 7
6-7
-

0~
3
I //1 ,_
FIRST CRACK OBS ERVED / / /5-6

N:
2
I

////(
a::
t- / / ,./4-5
(f)

2 [; ,//,t1 8-9-
w
=:! , / / I
(f)
2 I /
/ /
/ .
/ I/ I
W
I
/
t- / / / I

// .
I

I
I
/~/ ,/
//

.~
r:- /1'3-4
",/ ./~//
-.:;... ~~-::-~-:-;/ .... -
I

I
I
/
/

0
0
--==-
~~
0~
FLAT JACK
"---- ~
,-PRESSURE (k Po)
2
<t "'---
a:: "- \. I

1\9'-1~l2'3
10
t- f',
(f)
-1
w \\ I
>
(f)

~ -2 !

g: 0 100 200 300 338 400


:!:
o
u
Figure 30. Horizontal strains between reference points (Uriel and Perez, 1981).

5.0
0
Cl.
:::E
0-

W
0::
:::> 2.5
en
(f)
w
0::
Cl. Plate
W diameter:
l;i B=500mm
...J
Cl.
10 15
RELATIVE SETTLEMENT, 5/8 (%)

Figure 31. Tipical pressure-settlement curve for a plate bearing test (Valstad and
Str m, 1976).
187

Pressure cells

00 OC!
000
Abutment
o
o
Rockfill
00
o

Figure 32. Sketch of passive failure in situ test, devised by S. Uriel.


188

References

Bertram, G.E. (1973) 'Field tests for compacted rockfill' Embankment dam En-
gineering, Casagrande Volume, Wiley, N.Y, 1-19.
Bravo, G., Uriel, S. and Perez, J.R. (1988). 'Anti-cracking measures in the Cana-
les Dam', Our Work on Dam Construction, MOPU, Spain, 123-139.
Brinch Hansen, J. (1970) 'A revised and extended formula for bearing capacity'.
Danish Geotechnical Institute, Bull. No. 28,5-11.
Bureau of Redamation (1980) Manual de Tierras, Editorial Tecnica Bellisco,
Madrid.
Coba (1990) 'Rockfill embankments of the Torres Novas-Fatima section of the
northern motorway'.
De Beer, E.E. (1970) 'Experimental determination of the shape factors and the
bearing capacity factors of sand', Ge6thecnique, 20:4:387-441.
Del Campo, J. (1984) Pedraplen de ensayo. Presa de Martin Gonzalo, Unpu-
blished report.
Jain, S. and Gupta, RC. (1974) 'In situ shear test for rock fills', J. Geotech. Eng.,
ASCE,100:GT9:1031-1050.
Jimenez Salas, J.A and Justo, J. L. (1981) 'Geotecnia y Cimientos II'Rueda,
Madrid.
Justo, J.L. and Manzanares, J.L. (1983) 'Informe sobre las caracterfsticas de los
materiales utilizables en la presa de Martin Gonzalo y su repercusi6n en el dise-
no de la misma', Confederaci6n Hidrografica del Guadalquivir, C6rdoba (Espa-
na), Unpublished report.
Justo, J.L., Canete, P. and Del Campo, J. (1985) 'El empleo de rocas de baja
resistencia en los espaldones de presas de materiales sueltos', Revista de Obras
Publicas, Madrid, Mayo-Junio, 463-473.
Justo, J.L., Cafiete, P., Manzanares, J.L., del Campo, J., y de Porcellinis, P. (1988)
'The upstream facing of Martin Gonzalo rockfill dam', 17th Congo Large Dams,
San Francisco, 2:815-837.
Ranjan, G. and Prakash, S. (1975). Discussion to 'In situ shear test for rockfills',
J. Geotech. Eng., ASCE, 101:983-985.
Salas, J.AJ. and Justo, J.L. (1975) Geotecnia y Cimientos I, Rueda, Madrid.
Schmid, W.E. (1966) 'Field determination of permeability by the infiltration test'
Symposium on Permeability and Capillarity of Soils, ASTM, STP 417, 142-158.
Segovia, F. (1989) 'Estudio de presas de escollera con pantalla mediante un me-
todo de elementos finitos tridimensional' Ph. D. Thesis, Poly technical Universi-
ty of Madrid, Spain.
Sherard, J.L., Woodward, R.J., Gizienski, S.F. and Clevenger, W.A (1963) Earth
and Earth-Rock Dams, Wiley, N.Y.
189

Uriel, S. and Perez, J. R. (1981) 'Tensile behaviour of compacted clays by field


test', 10th ICSMFE, Stockholm, 2:581-584.
Uriel, S., and Perez, J.R. (1986) 'Previsi6n de la fisuraci6n transversal en terra-
plenes de carreteras y presas' Simposio sobre Terraplenes, Pedraplenes y otros
Rellenos, Soc. Esp. Mec. Sueio, Madrid, 607-625.
Valstad, T., and Str cm, E. (1975) 'Investigations of the mechanical properties of
rockfill for the Svartevann dam, using triaxial, oedometer and plate bearing
tests', NGI, 110:3-8.
190

Appendix 1: Infiltration from a shallow excavation

1.1. FALLING HEAD TEST.

Falling head test

x.
_ - - '_ _-'-_--L_..........:L-_---L_---.JL-_ _ _ _--.,I<.._-'::!.'Ia'=-'n~ot reference
dx.

Fig. Ap. 1.1. Falling head test in shallow excavation.

Hypothesis: flow is one-dimensional

(Ap.1.1)

Srw = degree of saturation after wetting


Sri = initial degree of saturation

Assuming the excavation is filled instanteneously, the solution of this equation is:

(Ap. 1.2)

dh dx
-n '" Sr ---- (Ap.1.3)
dt dt

Darcy's law:

Q dh h+x
v = ---- = - ---- = ki = k ------ (Ap.1.4)
S dt x
191

From (Ap. 1.2) to (Ap. 1.4):

dx ho n i\ Srx + x
n /I, Sr ---- = k -----------------
dt x

k x
------- dt = -------------------- dx
n/l,Sr h o +x(l-n/l,Sr)

k(l-n /l,Sr) x(l-n/l,Sr)


-------------- dt = ------------------- dx
n/l,Sr ho+x(l-n/l,Sr)

k(l-n/l, Sr) ho
-------------- dt = [1 - --------------------1dx
n/l,Sr h o +x(l-n/l,Sr)

k(l-n/l,Sr) ho h o +x(l-n/l,Sr)
--------------/I,t = x - ------------ In --------------------
n /I, Sr (1-n /I, Sr) ho

Substituting the equation Ap. 1.2:

I-n/l, Sr
h o- h----------
k(l-n /l,Sr) /I, h ho n/l, Sr
--------------/I,t = - -------- - ------------ In -----------------
n /l,Sr n/l,Sr (l-n/l,Sr) ho

1 /l,h 1 hon/l,Sr h o n 6 S r +(h o -h)(I-n/l,Sr)


k = -( ---------- ---- + ---- -------------- In----------------------------------)
I-n6S r 6t 6t (l-n6S r )2 hon6S r

1 6h 1 hon6S r ho-h(l-n6Sr)
k = -( ---------- ---- + ---- -------------- In ------------------)
I-n 6 S r 6t /l,t (l-n/l,Sr)2 hon6S r
192

1 Lh 1 hon LS r hn L S[1"l1
k = -( ---------- ---- + ---- -------------- In------------)
I-n6S r 6t I'll (1-n6S r )2 hon6S r

The hypothesis of instantaneous filling is difficult to reach.

Example:

Permeability tests in Martin Gonzalo test fill

P d = 2140 kg/m3 n = 0.20

Srw = 90%
I-n LS r = 1-0.2xO.8 = 0.84

ho = 25 cm

Graywacke, 6th layer:

6h = - 2cm 6t = 180 s

1 -2 1 25xO.16 23xO.16 + 2
k = -( ----- x ---- + ----- x ---------- In--------------)
0.84 180 180 0.84 2 25xO.16

k = 0,0132 - 0,0110 = 0,0022 = 2.2 x 10-3 cm/s

Lh = - 5 cm 6t = 360 s

5 1 25xO.16 2OXO.16 + 5
k = ------------ - ----- ---------- In --------------
0.84x360 360 0.84 2 25xO.16

k = 0,0165 - 0,0113 = 5.2 x 10-3 em/s


193

Slate, 3rd layer:

f',h = -11em f',t = 180 s

11 1 14xO.16+ 11
k = ------------ - ----- x 5.67 In---------------
0.84x180 100 4

k = 0,0728 - 0,0377 = 3.5 x 10-2 emls

Appendix 2: Infiltration from eased holes

Now let's see the result reached with equation 5 (section 5), with tube tests:

ro = 10.25 em
hI = 40 em
h2 = 20 em
t1 = 0
t2 = 192 s
f',Srn = 0.16

10.25 In2
kl = -------- -------------------------------- = 5.2 x 10-3 emls
4 3x20
192[--------------- + 1] 1/3
4xO.16x20.5

Equation 4 gives:

6,601.27
k2 = ----------------------- = 20.3xlO-3 emls
192x5.5x 1O.25x30
CHAPTER 8
LABORATOR Y TESTING AN D QUA LIT Y CONTROL OF ROCK FILL - GERMA N
PRACTICE
J. BRA UNS and K . KAST

1. Introduction

When looking at the poster prepared for this NATO - AS! on


"Rockfill Structures", one might be tempted to assume that
rockfills are granular masses of solid fragments of sound
rock.
As far as we have corne i nto contact with rock to be used for
dam fills at various sites in Germany and abroad, the nature
of the rock masses was more of the one to be seen in figure
1. In the major portion of cases, the rock was disintegrated
to a varying extent, resulting in a f i ll as shown in figures
2a and b.

Figure 1. View of quarry in disintegrated granite

'"
E. M arall/Itl das Nc\"cs (ed.) , Ad"onces in Rocl::[il/ Structllr /!s, L 9~219.
Cl ]99 1 KI""'er Acodemic Publishers.
196

Figure 2a. Fill of disintegrated granite (material placed)

Figure 2b. Fill of disintegrated granite (material compacted)


197

In the following presentation of our experience with testing


rockfill, we would like to point out the important aspects
of the breakdown in rockfills, because we believe that this
phenomenon governs the overall behaviour of rockfills in
most cases.

2. General Aspects

In 1959, our large scale triaxial cell for cylindrical


samples with h = 1.8 mjd = 1.0 m was put into use (see
figure 3).

Figure 3. View of large triaxial cell and sample


Soon after our first tests, we became aware of the effect of
particle breakage on the shear strength of coarse grained
materials. Even in well packed and apparently sound diluvial
fills most of the coarse pebbles underwent breakage during
testing at relatively low cell pressures. Figure 4 gives an
example of such material, a specimen (1 m in diameter) of a
material to be used for a dam in Greece.
At that time, fill dams planned in various regions of
Europe reached maximum heights of 150 m. In view of the
dimensions of these embankment dams, the effect of particle
198

strength on the overall strength of fills became of interest


to us.

Figure 4. 1 m dia specimen of well graded diluvial material

In several analytical investigations performed at our


institute then, the mechanical behaviour of granular masses
was studied by means of models like regular sphere packings
as shown - in the form of a representative element in
figure 5a.
We introduced the effect of critical contact force in
analyses of this kind (Brauns 1968) and were able to show
that the critical stress ratio in triaxial shear decreased
dramatically, if the lateral pressure exceeded a certain
value, which is defined by the strength of the particles
(see figure 5c). The packings used for the experiments were
made of glass spheres in order to create the brittle
behaviour of the grain material (figure 5b).
As can be seen from the diagram in figure 5c, the strength
of the mass is controlled by friction in the particle
contacts in the lower range of pressures, resulting in a
constant ultimate stress ratio (according to a constant
angle of internal friction). Beyond a certain cell pressure,
the bearing capacity of the sphere packing is exclusively
controlled by particle breakdown, resulting in an ultimate
stress ratio, which decreases rapidly with increasing
lateral pressure. Thus, for a non-randomly packed granular
mass, the stress regimes, where strength is controlled
199

either by friction and sliding or by breakage and collapse,


can clearly be separated from each other.

Oz:: a,

Figure 5a. Regular sphere packing as model for granular


masses (Element of model)

Figure 5b. Regular sphere packing as model for granular


masses (View of specimen for triaxial test)
200

FFkT~~'l
I,BREAKAGE
3,5

3
P
ct
2,5

2
~
1.5
1\I,-Jt49'fncrAL
0,8 8
1+L 80

Figure 5c. critical stress ratio in triaxial shear, theory


and results of tests (regular sphere packing)

Q7 35
MAX STRESS RATIO ...
'0>-----1

:\':L ~,;~, '


Q6 V
o LATERAl PRESSURE olllolN/m2)

.~_,
40

50

60-
81%1

Figure 6a. Behaviour of uniform quartz gravel (3 to 6 mm) in


triaxial tests (Shear strength (tan 0) and extent of
breakdown (B
201

When testing randomly packed granular masses, the well


defined borderline between frictional and breakdown beha-
viour effects disappears. This may be seen from figure 6
where the example of results of triaxial tests over a wide
range of cell pressures on a uniform quartz gravel (3 to
6 mm) with sound particles is shown. Figure 6a shows how the
maximum stress ratio decreases, and the extent of particle
breakage increases with the pressure level. The changes in
stress-strain and volumetric strain behaviour with increa-
sing stress level are shown in figure 6b. The higher the
stresses are, the larger the axial strains are to attain
maximum shear stresses and the more the densely packed mass
tends to undergo further volumetric compression during
shear.

SHEAR STRESS (01-03) IMN/m2)

.-----
12r---------------,

/03.SMN/m2

8 / "..0.2

6
/ /rI'.__.---.3

-2 " --.--
\'
"
6 12 18

--~.-_
24

-1,5
~

-4
\'
\\ "'.
.,
-6 '\
,' ""
\
-8
\ " " "'-,
-10
"", , " , '.3
- 12 .,
'-,
_~~----------'.~5~
'>UlUMETRIC STRAIN tv 1%1

Figure 6b. Behaviour of uniform quartz gravel in triaxial


test (stress-strain and volumetric strain relations)
What has all this to do with rockfill testing?
202

Figure 7 gives a comparison of the behaviour of three


different granular materials in triaxial shear under
relatively low cell pressures.

01 - 03 [kp/cm 2 ]
3o.---------------------T.~~lTTr.~~r.~~~C

SED

20

10

o Jlo;;;:=~.....::--+-----:7.t-5----i:l0----l~2.':"5---:115
'- :-------,...--
--- ----
E 1 ['1.]
-_ SED
5
G N7'-- _ _

10L-----------------------------~

EV ['/.] (COMPRESSION)

Figure 7. Behaviour of quartz gravel, well graded diluvial


sediment and coarse grained gneiss rockfill in triaxial tests
The thin solid line shows the behaviour of the uniform
quartz gravel (3 to 6 mm) mentioned already (compare
figure 6). The strength of the material is attained after
only a low percentage of axial strain and under dilatant
volumetric deformation (this means: mainly under sliding and
rearrangement of the particles). The thick solid lines
represent the behaviour of the well graded diluvial mater-
ial, of which figure 4 showed a large scale specimen.
As we can see, axial strains up to 10 % are required to
reach the ultimate shear strength here, and the volumetric
strains indicate compression with a slight tendency towards
dilation with large shear strains. Finally, the thick broken
lines show what was observed in the test on a uniform gneiss
material. (For the gradation of this material - and also of
the other two materials - before testing, see the gradation
diagram in figure 7). Even with axial strains up to 15 %,
the ultimate value of shear strength is not yet attained,
and the specimen was deformed with continuous volumetric
compression up to nearly EV = 10 % at the end of the test
(this is in spite of the tact that the material had been
well compacted with a vibrator plate in layers prior to
testing) .
203

This comparison clearly reveals that the behaviour of the


rockfill sample was predominantly controlled by breakdown
effects throughout the triaxial test.
This also becomes clear from the photograph shown in
figure 8a) and from the gradation curves given in
figure 8b). It is interesting to see (figure 8b) that even
the well graded river sediment undergoes severe breakdown,
when sheared under only 6 bar cell pressure.

Figure 8a. Degradation of coarse grained materials during


triaxial testing

SAND GRAVEL COBBLES


,
r- , ,
roo m m , -

W >-~
I ~'Jj
- 80
~

I
FLUVIAL SEDIMENTS
0, ~ 6 bar ~
~
iJ
'"
iii 60 ,,/' {1 - ..
3:
>- r--r-' ./' "'-/
ro I
40 i
;;"
- ,.- .L l ~L_-'-_

- - -- .- -
20 "..- V /~ QUARRIED
GNEISS
.-- .--' :.-- .--"" 11 0, ~ 4.S bar

a .- ~ f- '-'"
006 0.1 02 0.4 0.6 10 2.0 4 6 10 20 40 60 100 200 400 600
GRAIN SIZE [mm I -- ...

Figure 8b. Degradation of coarse grained materials during


triaxial testing (Gradation of gneiss and diluvial sediments
before and after testing)
204

What - in conclusion - one has to keep in mind, when talking


about properties of rockfill is that the mechanical
behaviour of such masses is mainly controlled by breakage
effects and that frictional effects are - almost - of no
importance at all. One may even doubt, whether it is
justified, to describe the shear strength of such materials
with a parameter like angle of "internal friction".

3. Aspects of laboratory testing

3.1 ROCK QUALITY AND GRADATION


The importance of breakage effects in the behaviour of
rockfill and the technical requirements in handling the
material in the laboratory as well as in the field in fact
largely depend on the "soundness" of the rock. To illustrate
this, we refer to a comparative study on granites with
different degrees of disintegration (Kast et al., 1985).
One problem which often arises is that a preliminary
design for an embankment dam must be based on the results of
core drillings only, that is: without any test quarries and
test fills. It is felt here that there is a need to work out
a methodology to quantify the degree of rock disintegration
on the basis of results of core drillings. It would, for
instance, be very useful if a method could be developed,
which makes it possible to predict - at least in a semi-
quantitative 'I'r,'} - the gradation of the fill and the degree
of disintegra~~on with the help of data collected from bore
holes.
We will not go into details of our comparative study on
granites here, but should mention that there are significant
relations between the mechanical properties such as strength
and modulus of deformation of core samples and particle
strength values on the one hand and the degree of chemical
and mechanical disintegration of the rock, on the other (for
details see Kast et al., 1985).
In fact, the best judgement of the properties of a
material and its suitability for a dam fill is to be gained
from a test quarry and a test embankment (compare figure 2).
3.2 "TRUE" ROCKFILL SAMPLES
The aim of laboratory testing of rockfill is to investigate
the strength and stress-strain characteristics in order to
predict the safety and the behaviour of the rockfill
structure under working conditions. As we have pointed out,
the behaviour of rockfill masses is mainly governed by
breakage effects. This holds for all phases after placement
of the material: the phase of compaction, and that of
loading under working conditions.
205

The actual inner structure of a rockfill embankment is


obtained by crushing down and "compactation" during compac-
tion with vibratory roller or falling mass. It is, there-
fore, not adequate - in our opinion - to excavate material
from a test field, bring it to the laboratory and recompact
it to the density measured on site. Not only is it difficult
to achieve the same dry densities as on site (with this
procedure), rearrangement of the particles will also not
allow to obtain what could be called a "true" or realistic
specimen with similar properties as compared with an embank-
ment, which has undergone "breakdown-compaction".
Figure 9 shows a comparison of test results obtained from
specimens with identical dry density, but which were compac-
ted by different means. One of the specimens was carefully
placed in layers and compacted by vibration, the other one
was prepared by just filling the material into a stiff
sample former and by compacting it by a falling mass. As can
be seen, the specimen prepared by the second procedure
reacts much more stiffly as regards the stress-strain as
well as the volumetric strain behaviour (compare Kast/Brauns
1981). It seems that this factor must be considered more
carefully, particularly if the parameters gained from such
testing are to be used to quantify stress-strain relations-
ships for numerical analyses.

1.0 DY~MIC COMPAC~'~ __ OP-~________ ~_._~

0.8 ().P .1) ,~..f:)' ~ GRANITE TYPE B


..0" /~ 0-30mm
0.6
Jf /'" VIBRATOR PLATE COMAtlCTION IN LAYERS

l"/
I ,.

0.4 u3 = 0.2 MN 1m2


0.2 {> / Qd = 1.77 g/cm 3

-1 12 E,[%]
-2
-3
-4
tv [%]

Figure 9. Effect of method of compaction on behaviour of


rockfill in triaxial shear
From similar experiences we have also have concluded that
"dry density" may not be a good parameter to define the
compactness of a rockfill mass.
206

In our opinion, there is need for further research on how


"true rockfill specimens" can be obtained for laboratory
testing.
For samples of small dimensions (20 cm in diameter), we
use the falling mass compactor shown in figure 10. The
material to be tested is filled into a stiff cylinder
resting on a heavy concrete block. The compaction energy is
applied by a falling mass through a transmissicn plate,
which covers the entire circular surface of the specimen. In
order to quantitatively describe the compaction energy, we
refer to the specific volumetric dynamic energy as defined
after Proctor for fine grained (cohesive) fill masses
(Kast/Brauns, 1980). Unfortunately, the falling mass method
cannot be applied to large scale samples due for obvious
reasons.

+---CRANE

HYDRAULIC
MOTOR FEED

~--- ~~~rSSION
i i r + - - - CYLINDER

r.~~~~~-.-~-~-~-~SAM~E
BASE

CONCRETE BLOCK

- ELASTIC SUf'PORT
/ 7 7 / ?

a) Scheme and section b) View of cylinder with


compactor
Figure 10. Falling mass compactor for preparing 20 cm dia
specimens (view)
207

3.3 SAMPLE DIMENSIONS


Another problem in providing realistic rockfill samples is
that of the necessary sample dimensions. Methods for
modelling rockfills or other ways to cope with the
coarseness of such materials are dealt with in other
contributions to this meeting.
In view of the fact that the specific particle strength
decreases significantly with increasing particle size, it
must be said that the results of laboratory tests are the
more reliable, the more similar the material is to that used
on site. This calls for large scale samples, because the
maximum allowable grain size in a specimen is restricted to
one tenth, or to one fifth - at least - of its minimum
dimension. A sample diameter or height of 2 m would thus be
desirable, but would also require immense technical set-ups.
A number of triaxial testing machines with 50 cm dia
specimens are available in various parts of the world today.
Our triaxial machine and a 1 m dia specimen has already been
shown (see figure 3). Materials from many places in the
world both far and near have been and still are investigated
in this machine.
3.4 LAYOUT OF DEVICES FOR OEDOMETER TESTS
Besides triaxial testing, the measurement of the com-
pressibility in oedometer tests is very common in rockfill
testing. The simplest method of testing is to use a stiff
cylinder of adequate dimensions. Figure 11 shows an
oedometer of this kind, 60 cm in diameter.

Figure 11. Oedometer test with solid stiff cylinder


208

As is well known, significant friction develops along the


side walls of this type of a cylinder when loading a
specimen (Hellweg, 1981), resulting in a non-uniform stress
state in the sample.
Special attempts have been made to prevent side friction
along the mantle of the specimen by using a cell made of
alternating stiff and soft rings. A cell of this kind is
available at LNEC/Lisbon. This design appears ingenious but
also causes slight technical problems. For example, intense
dynamic compaction in order to prepare breakdown-compacted
specimens is not possible in this type of device, for
obvious reasons.

____ - -----,-- I~.O I

OISTAt+C1E TUBES
TO AO .... PT '0"
, r - - - J_ ~ --, I
I
.
~
I
I
I
r- [--

IST_L~~~;:-I'o~o

l-r-r-r..t"auo.~ .~

"

: __L
Figure 12. Oedometer cell with stiff cylinder and special
load measuring device.
209

In designing a large scale oedometer cell for a research


institute in India (figure 12), we preferred the use of a
stiff cylinder with a smooth greased inner surface and a
thin protective cylindrical rubber sleeve, but also with a
special load measuring device: a circular load cell covering
only the inner portion (0: 60 cm) of the circular bottom
face (0: 100 cm) of the specimen. These arrangements make it
possible to base the evaluation of the test on the stress
measured in the centre portion of the cylindrical specimen
and, thus, to make allowances - to some extent - for the
side friction. Unfortunately, results obtained with this
set-up are not yet in our possession.
3.5 LAYOUT OF DEVICES FOR TRIAXIAL TESTS
Standard triaxial tests are normally performed using cylin-
drical specimens with h: d ~ 2: 1 and filter plates at
both end faces. Figure 13 gives a view of this type of
specimen after testing (50 cm dia rockfill specimen).

Figure 13. 50 cm dia Figure 14. h = d = 80 cm


rockfill specimen, rockfill specimen,
after testing after testing
210

As is well kown, barrel-type deformations, bulging or even


shear zones are normally to be observed with these testing
conditions. In one way or another the deformations of the
specimen are non-uniform and the evaluation of stress-strain
und volumetric strain relationships may seem difficult (Rowe
& Barden, 1964; Kirkpatrick et al., 1974).
In view of the recent developments in soil testing, the
question may be posed: Why not perform "element tests" with
uniform sample deformation through the provision of
- enlarged and lubricated end plates
- displacements of end plates without tilting
- h : d - ratio of 1 : 1
It seems that these testing conditions must be provided,
if we are interested in more than the shear strength only.
We have tried to construct a testing device of this kind,
for 20 cm dia specimens in a first attempt. positive
experience encouraged us to try the same for h = d = 80 cm
sample dimensions. In figure 14, a view of this type of
specimen after testing is given. The enlarged end plates and
the homogeneous sample deformations are clearly to be seen.
This homogenous deformation of a specimen can only be
achieved if the end plates are effectively lubricated.
Preparation of the smooth endplates with grease and a
protective rubber sleeve can be seen in figure 15.

Figure 15. Details of the new Figure 16. Removal of protec-


testing procedure with hid = tive metal strips after static
80 cm specimens: end plate compaction of the specimen in
with protective rubber sleeve stiff sampler former
211

As mentioned already, dynamic compaction with a falling mass


is not applicable for specimens of this size. Instead the
material is placed in two layers, each about 50 cm thick and
compacted by static compaction to the preset dry density
with the hydraulic ram of the triaxial cell. Very high
static loads of up to several hundreds of tons have to be
applied here which, in turn, calls for a very stiff and
strong sample former.
The rubber membrane inside the sample former would suffer
severe damage, if it were not protected by special means
during compaction of the material. A set of especially
shaped protective metal strips is used for this, covering
the entire cylindrical mantle of the specimen. Figure 16
shows how these metal strips are removed - one by one
after compaction and prior to placing the top plate and
sealing the rubber membrane against it.
According to our experience this testing technique makes
it possible to obtain very good and reproducible
experimental results.
Another aspect concerning the testing procedure should
briefly be mentioned and is illustrated in figure 17.

r-,
O. 40 load periods [min 1
N
,
E 0.35 r 100 1 '20 (5,'50
60 90 135 120 180
"-
z
~ 0.30
11 r 1 r
~~
(T)
l:l
0.25
lfl
I 0.20 ;.
l:l

lfl
O. 15
lfl
lfl
W
O. 10 ~
I
a:
I-
lfl
O. 05
> 0.00
w
0 0 2 3 4 5 6 7 8 9 10 11 12 13 14 15

AXIAL STRAIN EPS 1 [4J

Figure 17. Stress-strain relation obtained from a large


scale triaxial test with duration of loading steps.
In triaxial testing of granular soils the test procedure
with a constant rate of strain has become common. We have
found that rockfill should always be tested by applying load
212

steps. From these results it can clearly be seen that


rockfills need considerable time periods to react on any
change in stress. This is - again - mainly due to the fact
that any overall deformation of a rockfill sample is the
result of breakage effects (not of sliding effects), which
obviously requlre some time to develop. If we were to
perform triaxial tests with a constant rate of strain, we
would always overestimate the strength as well as the
stiffness of such materials under static load.
In closing this chapter, we would like to recommend and
encourage other investigators to apply testing techniques
similar to those described above and to report on their
experience for the sake of bringing rockfill testing up to a
level comparable to that, which has already been achieved in
soil testing for some time.
3.6. DIRECT SHEAR TESTS
To complete the picture, it should be mentioned that the
shear strength of rockfill (friction in shear planes) can
also be measured in large scale direct shear tests (shear
box tests). However, there are not many testing devices of
this kind in the world, as far as we know. An example of a
shear box apparatus is given in the following figure 18.

Figure 18. Sketch of the large scale direct shear apparatus


of the Central Soils and Materials Research station (New
Delhi) .
213

This machine has a shear box with a surface area 1 m x 1 m


and a height of 0.6 m. The working mechanism is similar to
that of conventional (small) devices in that the lower half
of the box moves, whilst the upper one is held in place by
means of load cells, which make it possible to measure the
shear force transmitted through the shear plane.
The vertical load is applied and kept constant by four
hydraulic jacks, which press against a stiff frame. With
this frame the device forms a testing unit, which is
independent of any additional loading arrangements, it only
requires the hydraulic pumps to operate the hydraulic jacks.
Thus, the device can easily be put onto a truck and
transported even to remote places to test materials on site.
Small size shear box machines of a modern design are
provided with special technical means, so that the upper
part of the shear box does not tilt during shear (Wernick,
1979). Provisions of this kind were not made in the design
of the large scale testing apparatus shown above, but
although this should have been done. Experience gained from
the use of the machine shows that tilting effects are
significant and require additional guiders to control the
adequate performance of the movable parts.
The stiff frame of the device, which is required to apply
the enormous vertical loads onto the sample surface, does
not allow the material to be placed in the shear box in the
testing position; a removal of the material after testing is
also not possible. In order to ease handling for the various
phases of the testing procedure, the device is designed so
that the entire shear box can be moved out of the loading
frame (sidewards) on special support beams.
CSMRS engineers are now gathering experience with this
testing device in the laboratory as well as under field
conditions. Improvements may be necessary for future
designs, as far as technical details are concerned at least.
It is, however, our opinion that testing devices of this
kind, which can be operated on site, may help to achieve
useful design data as regards the angle of shear resistance
of coarse grained fill materials prior to the availability
of results of elaborate laboratory testing programmes.

4. Aspects of quality control

The main purpose of quality control is to check on site,


whether the fill under construction meets the design para-
meters. In most cases - as far as we can see - the methods
for controlling quality are restricted to checking the
"density" of the fill by one method or the other, and to
checking the compressibility. sometimes permeability tests
are also performed.
214

In our opinion, the most important way to work out quality


standards is to use a test embankment and to carry out all
the necessary measurements in connection with this (figure
19). In such test embankments, we obtain - for instance
information as to how much breakdown of the quarried
material takes place during compaction (figure 20)_

Figure 19. Test embankment under construction.

SAND COBBLES
GRAVEL
BOULDERS
F
100 I
90
80 j:
70~
w
60s:
50>-
I,() m H+f.I+I-H-LH4l-I-i-H-IH+-

Figure 20. Gradation of quarried and compacted granite


rockfill.
In figure 20, a comparison of gradations is given. The solid
line on the right hand of the diagram gives the gradation of
a large sample taken from quarried material. The gradation
band shows what resulted from a number of sievings performed
on samples taken from the test embankment after compaction.
Sievings of large samples are necessary in such cases in
order to obtain realistic results. In this connection - by
the way - the problem arises as to how big a sample must be
in order to be representative, and at the same time not to
215

hide inhomogeneities. Performing gradation analyses for


rockfills is a cumbersome and time-consuming job. However
there is no way around it, because gradation is an important
factor for characterizing a fill, particularly in regard to
its permeability and to filter problems.
The next important parameter, which can be measured in a
test embankment, is the average density. The water replace-
ment method applied in connection with large test pits gives
good results here. Such test pits can also be used, from
time to time, during the construction of a rockfill embank-
ment, in order to control the compactness and homogeneity of
the fill.
Most valuable information can be drawn from settlement
measurements during the compaction of test fills (figure
21). From these observations, we can conclude which is the
best method to compact a certain material as regards the
permitted thickness of layer, number of roller passes
required, etc.

Figure 21. Settlement measurements during compaction ot a


test embankment.

Further information concerning the effectivity of a compac-


tion method can be drawn from systematic density measure-
ments in relatively small scale in test fills. If such
measurements are made step by step at various depths below
surface, the depth of influence of a compactor can be
calculated, as is illustrated with the results shown in
figure 22.
216

LANE 1, NUMBER OF ROLLER F'ASSES, n. 2

2.5 0
30
60
90
120
150

LANE 2, n=4
~--~~--~~~~~~~O
30
60
90
120
150

LANE 3, n =6
~--~~~~~~-r.~~-.O
30
60
90
120
1194 150
WIDTH OF FILL
(em]

Figure 22. Results of density measurements in a test


embankment
The dotted line in each of the diagrams gives the dry
density of the material, when simply placed and spread out
in 'a layer. The wedge shaped profiles show that the test
embankment was constructed with variable thicknesses. Each
of the wedge shaped profiles belongs to a lane of the
testfill with a certain number of roller passes: 2, 4 and 6
passes, respectively.
If we look into the data in some detail, we will recognise
that there is a SUbstantial scattering of the density
values. This is due to pronounced local changes in the
gradation of the material. Unfortunately there is no
reliable method to make corrections for changes in the
gradation, e.g. in the form of an analytical formula. Thus
the results of such density measurements can only be
evaluated and judged in connection with test embankments,
where a large number of such results are normally available.
This is why we do not believe that density measurements
are a suitable means for controlling the quality of compac-
tion - in isolated localities - during the construction of
a rockfill embankment. One should, in fact, not neglect or
belittle the psychological effect of density measurements
during construction, but in our opinion, the value of the
results of these tests does not stand in reasonable relation
to the extensive of time and work required to do such tests.
217

Dry density - in the light of quality control for embank-


ments is an indirect indicator. It is used to control
compactness and - through this compressibility. From
this point of view, it seems that the direct measurement of
compressibility is more adequate. A common means to test
compressibility are plate load tests, which can be performed
on test embankments as well as on embankments during
construction.

Figure 23. Special dead load arrangement for plate load


tests.
A disadvantage of such field tests is that heavy dead loads
are required (figure 23). The performance of such tests
takes a great deal of time. The equipment used for providing
the required dead weight is occupied for the testing period
and is, thus, not available for its proper use. This often
leads to the result that - in our region of the world
these tests have to be performed during weekends.
The duration of plate load tests is particularly long, if
saturation settlements are to be investigated under infil-
tration of water from the surface of the fill, since
deformations due to wetting of the material take their own
time.
The unfavourable relation between plate diameter and
maximum rockfill particles makes a reasonable interpretation
of test results difficult. We think that it is necessary to
control or investigate the nature of the material in the
zone just underneath the testing area by excavation of an
inspection hole after the test. If the test is performed
just above a massive piece of rock, the test results are
useless and cannot be taken into consideration.
Generally, we can only apply formulas describing the
settlements of circular footings, when evaluating plate load
tests. In cases where such tests are performed on relatively
218

thin layers of fill resting on a firm and rigid base, we


must not use the common formula for the settlement of
footing in soil of infinite depth, because the modulus of
deformation of the fill would then be overestimated. In such
cases we may use the formula given in figure 24, with a
factor a to be chosen according to the local conditions of
relative layer thickness.

(1-
PROBLEM: 0,1 0,2 0,3 0,4 0,5
00

t
3EI
4
I 2r I 6
z
8
lEvi
10
E; 00 12
z/r 14
MOOULUS OF 16
DEFORMATION : 18
Ap 20
EV= a. rAS

Figure 24. Evaluation of plate load test on layer of finite


thickness.
In closing this chapter, we would like to emphasize once
more the importance of plate load tests for direct quality
control of compacted embankments including rockfills and to
point to the difficulties in the performance and interpreta-
tion of density measurements.

5. Concluding remarks

The intention of this contribution was to point out a number


of special factors in the behaviour of rockfill. It seems
important to keep in mind that rockfills vary in nature and
behaviour within wide borderlines. Weakness of materials
plays a most important role, but even in sound rock,
breakage effects are dominant. This brings up the problem of
how to properly handle the material for testing.
Besides the preparation of realistic specimens, conditions
of homogeneous deformations in triaxial testing seem to be
important, if stress-strain behaviour is of interest in
connection with numerical analyses.
The prediction of the nature of the rockfill the from
bored samples during the early stages of investigation needs
219

more attention, because as far as we can see


experience gained from recent test quarries and test embank-
ments have often revealed unexpected behaviour and
surprises, leading to changes in the design of dams.
As regards quality control, more emphasis should be given
to testing the deformability of fill than to density
measurements. Unfortunately, we do not have any idea as how
to make tests on deformability easier or more effective than
they have been in practice up to now.

References

Brauns, J. (1968): tiber den EinfluB des Einzelkornbruches


auf die Belastbarkeit von Haufwerken, besonders von
regelmaBigen Kugelpackungen, im Dreiaxialversuch. Ver-
offentlichungen des Institutes fur Boden- und Felsme-
chanik der Universitat Karlsruhe, Heft 33, 1968.
Brauns, J.; Kast, K. and Blinde, A. (1980): Compaction
Effects on the Mechanical and saturation Behaviour of
Disintegrated Rockfill. Proc. Intern. Conf. on Compac-
tion, Paris, Vol. I, pp.107-122.
Hellweg, V. (1981): Ein Vorschlag zur Abschatzung des
Setzungs- und Sackungsverhaltens nichtbindiger Boden
bei Durchnassung, Mitt. Inst. fur Grundbau, Bodenme-
chanik und Energiewasserbau, Universitat Hannover, Heft
17.
Kast, K.; Blinde, A.; Brauns, J. (1985): Verdichtungs-
Verformungs- und Sattigungsverhalten von Schuttungen in
Abhangigkeit von der geologischen Gesteinsentfestigung.
Ingenieurgeologische Probleme im Grenzbereich zwischen
Locker- und Festgesteinen. Hrsg.: K.-H. Heitfeld.
Springer Verlag, Berlin/Heidelberg, S. 237-254.
Kast, K. & Brauns, J. (1985): Influence of the Extent of
Geological Disintegration in the Behaviour of Rockfill.
Proc. XI. Int. Conf. on Soil Mech. and Found.Eng., San
Francisco, Balkema Rotterdam, Vol. 4, pp. 2359 -2362.
Kast, K. & Brauns, J. (1981): Dynamic Compaction of
Rockfill Samples. Proc. X. Int. Conf. on Soil Mech.
and Found. Eng. Stockholm, Balkema Rotterdam 1981. Vol.
1. pp. 669-671.
Kirkpatrick, W.M., Seals, R.K. and Newman, F.B. (1974):
Stress Distributions in Triaxial Compression Samples.
ASCE, Vol. 100, No GTZ, pp. 190-196.
Rowe, P.W. & Barden, L. (1964): The Importance of Free-
ends in the Triaxial Test. PASCE, 8M 1, pp. 1-27.
Wernick, E. (1979): Bestimmung von Bodenkennwerten mit
einem parallelgefuhrten direkten Schergerat. Die Bau-
technik 9, S. 307-313.
CHAPTER 9
CREEP OF ROCKFILL
A K . PARKIN

I. Introduction

In gCOIcchnical engineering, lime-dependent settlement is nonnally associated with the process


of consolidation. In \his, settlement behaviour is dClennined by the TatC at whic h water is able
10 now from the voids under a hydraulic gradient. allowing particles to slide into a more
compact arrangement.
Rocklill di ffers from consolidating soils insofar as it is coarse, and therefore free -d rai ning,
and in being .mgular, such thaI particle fragmentation under high contact loadings accounts for
a significant wmponcnl of comprcssivc dcfonnation. Therefore. in an ocdcomclcr compression
\eSI, any hydrodynamic effects will not extend beyond some minutes at most, so that they arc,
in praclicallcnns. not observable. The SClllemenHime relationship. in semi-logarithmic fonnat
( Fi g. I) thcn consists of a very rap id initial compression, accounting for perhaps 70 to 80% of
the tow! 24 hour compression. foll o wed by a linear section whose slope may be described by
the coerficicnt of secondary compression e" (Wahls. 1962)1

(I)

T his phase of the settlement process is known as creep, and is essentiall y similar to
secondary compression in clays (where the mechanism ma y, however, be very differem).
The analysis of rockfill creep will not nonnaUy be a major consideration in dam design.
Generally a load-defonnation graph (or e v. log p) prepared from 24 hour selliements in a siage
load test will give adequa te pred ictions of selliements in situ. To this. a designer may choose
to add a nominal margin (say 5%) to allo w for creep. in order 10 ensure that Ihe cres\. fo r
example. ha~ ~urficicnt camber to maintain the requ ired freeboard. However, Ihe analysis of
creep becorno:~; muc h more imponant fo r the diagnostic analysis of settlement records. or in
situations when:: it is necessary to make long-tenn predictions of creep selliement. The success

I. The nature ofC" has ocen studied by Mesri and Godlewski ( 1977) who showed. for cohesive wits.
Ihill C~ is strongly depemlem on effective stress and thm the ratio C,)C, fall~ witllin fairly narrow
limits. It is. ho,,eveT. yet to be established whether such results also relate to cohesiOllless
materials. bill il is of inlcrest thatlhcy note thaI C" can change wilh lime and speculate on foctors
that mlly emtse this.

22J
E. Maran/Iil lias Neves (ed.). Adwmcts in Rockfill SlrUClure~ . 221-237.
C 199t Klu ...tr Academic P"bli~hers.
222

of such analysis then depends on an understanding of the fundamental nature of creep, and not
merely the application of conventional soil mechanics procedures.

Time, mins.
01r---,----,r-rr..,...,mlTOO.::......._.-..,-..,-.,-r-rnlOOO;:r::::::................,
I I I I I II II I I I I I I I I I

250 kPa

~
E
10F~~-------- __- ______~500~~kP~a~_
c
CI>
E
~ f-
a;
VI

2Fr-~~~~ON~----------------~10~0~0~k~p~a
3A MUDSTONE __
BLUE ROCK DAM
~d - 21 t/m 3

Fig. I Deformation-time graph for mudstone rockfill (Parkin, 1983)

2. Rate Methods Applied To Settlement Analysis

Since the origin of consolidation theory in the 1920's, settlement analysis has almost invariably
focussed on the settlement-time relationship, whether plotted to natural scales or some
convenient mathematical scale, as in the traditional methods of evaluating coefficient of
consolidation, Cv ' Recent work, however, has shown that a velocity-time graph, produced by
differencing raw settlement observations, is not only fully sufficient for the evaluation of
cv , but yields a significantly more accurate value through the elimination of irrelevant inputs
(Parkin, 1978; Parkin 1981).
This work forms the basis of the Velocity Method, whereby Cv is determined from the direct
overlay of a theoretical U-T solution on an experimental velocity-time graph (Fig. 2). In this
procedure, no geometrical constructions ale required because the settlement rate is everywhere
quite independent of initial conditions, and, over the matched region, also independent of
secondary consolidation. By contrast, traditional procedures (tw, etc) require a consolidation
state to be established in relation to initial and final states, both of which arc irrelevant and
neither of which can be established without ambiguity.
In the post-primary region, the settlement rate curve for a clay usually becomes linear (rather
than following the continuously curved theoretical solution) at a slope Cp which varies with
load. As indicated in Fig. 3, the decay rate, Cp, passes through a minimum for a load
increment straddling the pre-consolidation pressure, Pc (corresponding with the increase in C"
reponed by Mesri and Godlewski, 1977). The reason for this variation is not clear, but it may
relate to variations in the time of origin of creep. However, for coarse permeable materials
223

10r-----------,----------,---,

01

c:
E
"- ....
E

~
.::>
E
'"
" "
'"

.
a:
c: a: 001
.12

~\
"
"E
"!1 '"
E
~ 01
2.-""T/.
.!!
c:
0
u
U~ Q;
(/)

0001

0'01 L -________- - '_________- - '__- '


001 01 0-1 10
\
Time Factor T Time t (min)

Fig_ 2 Terzaghi solution and oedometer consolidation. 2 (after Parkin, 1981)

such as rock fill , a primary region will not be observed and the settlement rate curve then
consists of a straight line with C~ = 1, creep commencing effectively at the instant of loading.
This value of C~ = 1 occurs for transient creep in a wide variety of materials, and appears
to be best expiai",;d in terms of the stochastic theory of Pusch and Feltham (1980), as the
altemative r81e process theory is limited to the prediction of steady-state creep rates only
(Mitchell et .11.. 1968). Because C~ is mostly ncar unity, it follows that the integral
relationship, plaited as settlement against log time, will also be in general (but not always)
linear, although with varying slope Ca- This, the traditional basis of creep analysis (as used
in procedures ~;uch as Poulos et a1. 1976), would be the simpler one to use, except for the fact
that C a is not casily defincd from laboratory tests 3 , and often exhibits erratic variations, as in
one case of reponed crecp settlements of somc major foundations on a mudstone (Mcigh, 1976:
Fig. 4). Factors which can contribute to such behaviour will be discussed subscquently. It
may be noted, however, that whereas C a is a material propeny derived from testing and
dependent on effective stress, C~ is essentially independent of any material propeny, except
as it occurs following consolidation.

2. The initial point on this graph is high because it includes clastic compression, and the subsequent
departures are the typical consequence of a sticking dial gauge. Note, however, that such
interpretations are not possible on the usual settlement-time graph.

3. Sowers et a!. (1965) repon values of ea from laboratory tests on rockfill that are generally about
half the field value.
224

r I
LOADING (k Pa)

o 25 (seating)
50
.100
0200
m 400
6800

10 100 1000

Time (mins)

1.8
I
Johnson 5t. Bridge
BH.201
rl
I
1.6 t Sample 271 gil C V

~
1.'

~
-- t-H-
:
~~-~<
' I I

/
1.2
a: I
..... ~
I

1.0
I
~
C
0.8 ~ 1\ i
I !

0.6
10
I It 100 1000
Pressure (kPa)

Fig. 3 Effect of Pc on settlement rate of a silty clay (after Parkin, 1981).

3. Rockfill Creep In Oedometer Compression

Rocklill creep can be studied in the laboratory in an oedometer test, wherein the load
application is quick so that the time of origin of creep, to' is clearly defined (consolidation
being insignificant). Should to be not known, as is often the case for field records or for creep
developing after consolidation, then the slope of the creep line on a velocity plot may not be
physically meaningfuL In addition, laboratory observations will usually be more consistent
than field measurements because of greater control and less disturbance from various sources.
For any load stage, a velocity diagram can be prepared by differencing the raw settlement
readings. Typical is a result for a soft siltstone rockfill from Sugarloaf Dam, Victoria (Fig. 5),
which shows firstly a slope that is close to -1, as it invariably seems to be, and secondly a
sudden discontinuity. Whilst the occurrence of such discontinuities is not predictable, it is
notable that the slope of the creep line is invariably preserved across them. Clearly, should
225

such a result be presented on a conventional settlement-log time graph, a discontinuity of slope


must occur. Any attempt to fit a single curve in such circumstances is tantamount to fitting
a single curv~ through the data of Fig. 5.

o
S!

.
c
E
~
:::0
~ ~ f----+-+-+--"4r+--W-/~

Time, t, months (t=O at 1 1 64) Time, t, months (t =0 at 15464)

Fig. 4 OJdbury A - Settlement of reactors R 1 and R2 (after Meigh, 1976)

Discontinuities of various forms (of which Fig. 5 is one example) prove to be an intrinsic
feature of creep almost wherever it occurs. Similar effects have been observed on clays,
rockfills and soft rocks and described as "limited instabilities" by Bishop (1974), based on the
work of Bishop and Lovenbury (1969) and Pigeon (1969). In particular, Bishop noted that it
was not possible to eliminate these random features, even with the most meticulous care. Their
origin would appear to lie in the stick-slip nature of creep which distinguishes it from the
fundamental continuity of the consolidation mechanism: inherently, as load is transmitted
through a material, local stress concentrations develop, followed by yield, in diminishing steps.
In the case of Fig. 5, the discontinuity represents a locking-up of the system, which may relate
to the development of normal or shear stresses on the oedometer wall.

4. Application To Field Settlement Records

If the vclocity method is to be applied to the analysis of field settlement records, there arc
generally two problems. The first is that the time of origin of creep to may not be readily
identifiable, and the second is that field observations may contain an appreciable component
of scatter.
226

.0 1
I I
Random Fill
.005 ~ Sample No.3--
"'- p=1.94t1m'

.002 ~ cry = 770 kP

~
-
.5 .001
E "\.
~

1!"
.0005

.0002
""'" "
t
o .0001
"-

~
.00005

.00002
~
50 100 200 500 1000 2000 5000
Time (minute.)

Fig. 5 Creep rate from ocdometer test, Sugarloaf siltstone (Parkin, 1985)

One possible means of estimating to is to plot the reciprocal of creep rate liS against time,
whereupon, if a sufficiently good line results, to can be read as the time intercept (Fig. 6). This
example relates to Cedar Cliff Dam, a sloping core rockfiIl dam in North Carolina for which
a settlement record (S v log t) is given by Sowers et aI. (1965)4. The accuracy with which
such creep rates can be calculated is limited by the accuracy with which a settlement graph can
be read, and it is clearly preferable to have access to original records for this purpose.
However, this analysis would suggest that to is rather later than middle of construction (April,
1952) as plotted by Sowers et aI., probably because of a significant contribution to creep
coming from the subsequent water loading on the acutcly-inclined core (the dam having filled
soon afler c(Jmpletion in August, 1952). With this value of to' the velocity plot becomes
closely linear at a gradient of -1, as is generally found to be the case.
Where th(']'( is a measure of scatter in field data, (as, for example, Fig. 4), that scatter will
be considerably amplitled on a derived velocity-time graph. It is important then to understand
that the velocity graph contains no additional error whatsoever and that the original data can
be recovered precisely by integration. Furthermore, whilst the extrapolation of a velocity line
for the prediction of creep settlement may on occasions seem imprecise, it will never he any
less accurate than any other method.

4. A description of Cedar Cliff Darn is given by Growden (1958), but docs not include the settlement
records quoted by Sowers ct aI. (1965).
227

On some occasions, it may be found that the scatter of data is of such magnitude as to make
the velocity plot of little value. Such scatter will tend to be biased below the mean creep line
because of the logarithmic scale, and may even include negative settlement rates. In such
circumstances, it may be found that a useful trend emerges if settlement rates are computed
over a double, or even triple time increment, at the penalty of some loss of sensitivity.
The value of a creep rate plot then depends on whether or not it is possible to identify and
isolate discontinuities. Some possible types of discontinuity are examined by Parkin (1971),
such as Fig. 7, from which it can be seen that, in the presence of a discontinuity, a
conventional settlement-time graph consists of two quite unrelated segments. Whilst these
segments might be clearly visible as drawn in Fig. 7, such curves normally have to be drawn
through plotted data incorporating varying degrees of scatter. There is then no way of fitting
an accurate mathematical funetion to the data that is in any way reliable for the prediction of
creep scttlemcnt. Normal practice in this case would be to fit a single continuous curve, which
must clearly be too steep, leading to an over-prediction of settlement.

150

A
V
50
V
~
V t o- 4
y"'"
.c-d
200 400 600 600 1000
lIS months I percent

...
1"-"-
0.03

0.02
:; i'-
c:
o
E
f',~
r--
'E 0.01

.
"~
~
'"'"
c.

-m 0.005

"
'OJ
a: ~

.
'"
_ 0.003

.
c:
E
.!! 0002 ~
\1 "-
'" "r--
,
m

10 20 30 50 100
(t tol months

Fig. 6 Dete mination of to' Cedar Cliff Dam, Parkin (1977) from data from Sowers et a!.
( 196:;\
228

Because the slope of a creep rate line is not a material property, but is unique to the
mechanism of creep (and the value of to)' the slope cannot change across a discontinuity, as
seen in Fig. 5. This may be quite important, as a discontinuity can occur at any stage and may
not always have an adequatc span of data following it to define a slope. Furthermore, the
scatter in thcse points will increase because they are derived from increasingly small
movements. Whilst it is rare to observe more than one discontinuity in any creep record, that
possibility aiways exists. However, unless the structure is approaching collapse, any
subsequent discontinuity will be of diminished magnitude because the creeping system is
always moving towards an optimal distribution of stress.
It is often unfortunate that the evaluation of dam settlement records is limited on account of
settlement observations not being commenced soon enough. This occurs because the
installation of crest settlement monuments is often delayed until all operations likely to interfere
with them have been completed. Whilst this is indeed important in respect of total settlements,
it is altogether irrelevant to readings of settlement rate, where loss or damagc to a reference
point is only of transient significance to the creep rate record. The importance of providing
temporary reference marks at an early stage should therefore not be overlooked.

t --+ log t

s
1 ""-
"
Fig. 7 Creep discontinuity caused by stress change (after Parkin, 197 I)

S. Crest Settlement of Dams

The velocity method has been applied to a number of rockfill dams under the control of the
Hydro-Electric Commission of Tasmania (HEC), in order to examine settlement rate behaviour
and what, if anything, can be deduced from it. These results, made possible only through
having access to original survey data 5 , have been presented previously by Parkin (1985), and
are reproduced here.

5. As indicated earlier, settlement-time graphs cannot normally be read with sufficient accuracy to
allow a meaningful velocity analysis to be made, unless settlements are large, as in the early
dumped rockfills.
229

Of the many structures with available data, five were selected for analysis, as listed in Table
1. In other C(ises, the time base was found to be too short, movements too small, or the
behaviour irregular for some reason, such as the leakage history at Scotts Peak Dam (Cole and
Fane, 1979). For each of these dams, settlement rate diagrams were prepared from precise
levelling records on crest marks in the vicinity of the maximum section.
In addition to settlement readings, settlement rate studies are also dependent on the chosen
time of origin for plotting. Arguments can be advanced for taking this at the mid-point of
construction or at the end-of-construction (E.O.C.), but all construction programmes differ and
even such dates as these can be difficult to pinpoint, in contrast with the laboratory situation.
In all the following graphs, E.O.C. of the fill has been used as the estimated time of origin, but
because of these uncertainties, the gradient of the creep line may not conform to -I in all cases.
For a time origin based on E.O.C., it would appear that crest settlement rates satisfactorily
conform to a straight line relationship, at a slope that is normally close to -1. This is best seen
in the result for Serpentine Dam (Fig. 8), wherein the embankment consists of rolled weathered
quartzite and schist, over some 8 m of in-situ river gravels (H.E.C., 1975). E.O.C. has been
taken as September 1971, when the embankment was complete except for a final l.2 m of
rockfill added after completion of the concrete facing. The significant features of this graph
are, firstly, that the application of water load during reservoir filling (which occupied a
considerable period because of the very large capacity) appears to have no effect on vertical
settlement, a characteristic that applies generally to faced rockfills, and, secondly, that the
scatter of points increases with time as the movements to be measured become very small.

Time from E.O.C. (months)


5 10 20 50 100 200
5
~
Storage filling
-
.c
c:
0
E
2

"E
E

~
0
0:: 0'5
0.
CI>
~
U
02
Chainage 234

01

Fig. 8. Settlement Rate, Serpentine Dam


230

Table I. H.E C. Dams in Tasmania

Dam Location Height (m) Type

Wilmot Wilmot R. 30 Faced


Parangana Mersey R. 55 Zoned
Rowallan Mersey R. 42 Zoned
Cethana Forth R. 110 Faced
Serpentine Serpentine R. 30 Faced

Wilmot Dam, of interest in being the first faced rockfill to be built by the H.E.C. (Cole,
1971), is ostensibly similar to Serpentine with respect to its creep behaviour (Fig. 9). In this
case. the rocUilI is a hard greywacke, placement of which was completed in November 1968.
Wilmot, however, is unique amongst this group of case studies in showing a gradient rather
steeper than -1, which suggests that the E.O.C. date may be inappropriate for some reason. A
closer scrutiny of operations in the 16 month period up to the commencement of crest
settlement readings shows that face construction took a further 10 months, during which a
settlement of lOmm was recorded from hydrostatic settlement gauges. Negligible settlement
occurred in the succeeding four months up to the commencement of filling, indicating
behaviour that is not at all consistent with the line drawn on Fig. 9. The indications are
therefore that a new phase of settlement has commenced during filling, and a time of origin
taken at this point does indeed lead to a line of slope -1.

Time from E.G.C. (months)


2';.o_---.:2~0~-----'5~0~---..:I:::;.OO~____,


05
""co -J1~
E
......
E
Storage
filling

S 02

III

"
tl:
a.
III
01

III

U 0.05

Choinoge 320

0.02 L...._--L_ _ _L-_-L_-----l

Fig. 9 Settlement Rate, Wilmot Dam


231

As compared with faced rockfills, cored dams show notable differences in their settlement
rate behaviour. A typical example is found in Parangana Dam (Mitchell et aI., 1968), where
the central core, containing the crest marks, consists of weathered granodiorite, supported by
rockfill shoulders of quartzite and schist. Based on the embankment E,O,C. at June, 1968, the
creep rates define a 45 line over the greater part of the time spanned, except for some higher
rates in the initial stages (Fig. 10). These are evidently attributable, as indicated, to reservoir
filling, with settlement associated with the penetration of a wetting front, and, dl a later stage,
to either a 5 month period of reservoir drawdown or (more probably) to road paving operations
on the crest. Such variations appear to be superimposed on an otherwise global trend for the
main body of the dam.

Time from E.O.C. (months)


10 20 50 100
10
........ Storage filling

..,-CreSf p<:IVino

~
"- '\.
8E "-
"-
E
5
"
"..
05
n:
a.
l'
u
0'2

01
Choinoge 400

0'05

Fig. 10 Settlement Rate, Parangana Dam

In yet other cases, the interpretation of settlement rate behaviour is made difficult by an
excessive level of scatter, which, in itself, must exist for some reason. Such is the case of
Rowallan Dam, a cored dam with a central concrete spillway, wherein the core consists of a
well-graded weathered till of quartzitic and dolcritic origin, and the rockfill shoulders arc of
quartzite and schist (Mitchell, et aI., 1968). For E.O.C. at January, 1967, the creep rate
behaviour at maximum section is as depicted in Fig. II. This shows a clear global trend, at
a slope ncar - I, "lith a substantial superimposed scatter, the magnitude of which is too great
to be explained by rounding errors in the surveyed levels. Filling was gradual, with the final
third occurring during months 7 and 8, where the creep rates reflect the passage of the wetting
front. A low rOini around month 23 may be connected with a reservoir drawdown caused by
pi ping problems (Mitchell ct aI., 1979), but the fluctuations around month 40 arc for reasons
not ascertainablc; by the Author. A correlation of creep rate against rainfall has also been
attempted and gives some grounds for regarding the scatter as a seasonal shrink-swell
phenomenon in the sometimes rather plastic core.
232

Time from E.O.C. (months)


1 2 5 10 20 50 100
10r----.------.-__ .----,------,----,r---,
.~

5
....--.-..
Storage filling
:c
co
E
"-
E 2
.

1"
o
a::

c.
0>
~ 0-5
u

02
Choinoge 1400

01L----L------~---L----~----~----~--~

Fig. II Settlement Rate, Rowallan Dam

As with Rowallan, Cethana Dam also shows substantial scatter in its creep rate behaviour
(Fig. 12), but still with a global trend at a gradient close to -1. This structure is a well rolled
quartzite fill, placement of which was essentially complete by November 1969, except for a
further II.5m added in October 1970 (adopted E.O.C.) after construction of the concrete face
(Fitzpatrick et aI., 1973). In this case, however, the crest level marks are located in the parapct
wall, so that recordings thereon are influenced by the movement of the facing slab. This is
reflected dramatically in the high settlement rates on first filling, rising to a peak of 20
mm/month (off scale) in April 1971, as the lake rose 30 m. This is clearly concrete shrinkage
in contact with stored water at a temperature typically around 6C. Thereafter, high and low
creep rates mostly correspond to a rise or fall in reservoir level during the corresponding time
interval, confilming the temperature influence, while the underlying rockfill continues to creep
independently, dominated by the final lift, and such that only a resolved component (COS2~) is
recorded on the parapet wall.

6. Load Tests on Large Bored Piles

Some additional case histories of creep can be drawn [rom an investigation into the behaviour
of bored pilcs sockcted into weak rock, carried out by Williams (1980). In this investigation,
a series of test IJiles were constructed in a highly to moderately weathered siltstone from
essentially the same sequence as the rockfill of Fig. 5. The sites were located in a motorway
cut of some oim depth (S) and a brick pit excavated some 25m into rock (M), both near
Monash UniV(T;ity. These locations allowed the tests to be taken to failure and excavated
afterwards for .:xamination.
233

Time from E.O.C. (months)


I 2 5 10 20 50 100
5r---~------r---~----.------'-----.----'


2
.;:
c:
0
E
"-
E
E

0'5
.. ~
II> Storage filling
0
0::
a.
II> Water level rising
~ 02
u Water level falling f

01
BLOCK H

005 L -_ _---'-_ _ _ _ _ _. L -_ _--'-_ _ _ _-L-_ _ _ _---L_ _ _ _- ' -_ _- - - '

Fig. 12 Settlement Rate, Cethana Dam

Some of these tests were designed to operate in side friction only (with a collapsible base)
and some to operate in end bearing only, as documented by Williams. Loading was applied
in stages, generally of one hour duration, and increased up to failure, with sufficient
information available for the construction of velocity-time plots in many cases.
Results from two end bearing pile tests are shown in Fig. 13, both as raw plotted data and
with an interpretation superimposed. In the case of test M5, the loading is well below failure
for this location (F = 2.8), and the creep is of a routine character, following a 45 line. In the
case of test S4 (on a softer rock), some more complex behaviour is apparent, but a coherent
picture emerges if a 45" set square is used to construct segments of 45 lines, as on the right
of the figure. It is then apparent that there is a development from a stable creep at a load of
3000kN into a rapidly deteriorating condition at 4350kN, evidently with the successive
formation of cracks. In fact, some tilting of this pile commenced at 2550 kN, increasing
significantly at 3500 kN, and increasing further to 1 in 27 at the maximum load, at which point
the pile was deemed to have failed (load increment 3900 kN not ploned). Being a failure
situation with a potential for collapse, these successive slips do not show any tendency towards
diminishing magnitude.
A typical result for a side friction pile is shown in Fig. 14, again presented as both raw data
and with an interpretation superimposed. Load was increased at approximately 500 kN per
stage up to 4900 kN as plotted, where creep settlements became substantial. The load was
subsequently taken briefly to 5100 kN, but settlements were too great and the pile was deemed
to have failed. In this case also, a 45 set square can be used to fit segments of 45" lines, in
which a confusing set of points is seen to define a series of small and diminishing slips. Final
excavation showed that shearing had occurred generally through the roots of the deliberately
formed helical a:,perities (= 12mm depth) in the socket wall, creating a shear zone up to
100mm thick.
234

1.0
+
TeST S4 -
0; 1.00m - : = j :~ ::i
0 0 L fD ; 0
-
.- - _ .
0
0 02
0 0 8
a = 4350 kN +
0
orF'b
c: a = 3500 kN 0 0
E a; 3000 kN 0.'

-=
.. +

~ 0,2
'";;;
Q) 0 0

c: 0. '
il +

. .
0.02 - - "'-",------1, .,
~-.; + +++
U)
0." 0.001 : I ---------~.

.. ..
~-----------~

0,02 TEST 5M TES' 51.1


D = 300 mm D; 300mm
0.0' _ LID = 3,3
L. = 3.3
'D
a = SOOkN a = SIX) f.N
-------__---,--------0--
0 ,00< L--_ _ __ _ _ _' - -_ __ _ _ _- - "

'0 20 '00

Time (mins)

Fig, 13 Creep behaviour in end bearing piles

7. Conclusions

The senlement rate method has been proposed as a means of monitoring and interpreting creep
settlements, with examples drawn from crest settlement observations on a number of rockfill
dams. It is shown that, provided an adequate time origin can be identified, the creep rates
generally follow a line of slope -Ion a log-log plot, subject to transient deviations that can be
identified and related to events in the construction and loading history. The reason that
interpretations are possible is that, as for consolidation, the effects of these events can be
viewed in isolation from all that precedes or follows. This is not the case when data is
examined in the conventional format of a settlement-log time graph, wherein much potentially
useful information can pass unnoticed. Thus the velocity plot has an important diagnostic
function for creep, as anticipated by Push and Feltham (1980),
Rate analysis may not always yield significant new information, and the results herein cover
a spectrum. Interpretations are in some cases conjectural, and in others information may not
be available or the behaviour may be too complex for interpretation. The potential of the
analYSis is, however, enhanced if data can be obtained early in the life of a structure, from
temporary marks if necessary, but normally Authorities prefer to delay the establishment of
permanent marks until construction activity has ceased. The fact that E.O.C. for a dam seems
to provide an appropriate time of origin for creep might suggest that overall creep settlements
235
1.0

0
- ~-;t-=b: :1-:::. I .
TESTS3
0 = 1m.
0.4
UD = 2.5
C I-+-+--t - f--t--- t- , - - Q=4900 kN
~... 0.2 ._-- . .-- r '
~'0 0"
I- +-+--+--HI--- -- --- -----.
*
0.' .::> ---------
, It') ('I

a:: 1-4--1--1--1-+- - - - - --- .-.- -- 1-' , , --- -


c: "
Q)
\ - - f---t----I-I- --.-1----- -- -.
E
Q)
EQ) 0.04 1-+--1---+-+-1-- - -- -- -- -- -_Q..- -_.
.)
III I--I--\--I--I-I---I--t-f--- - - _. ---.-- - -- -- ,-,
0 .02 I-+-+--I--I-II---j_. - -- - - - - - - --. . .

0.0'
2 10 20 40 100 200 400

Time (mins)

1.0
-'f,,.-t--I--t-t-- -I- - - -. --l- - J:.:1--
1"\ _- - r -
"<
_ . -- TeST S3
i' _ ___ D =1m.
0.'
.:
~... 0.2
__ ____ :- ['s~I~ -- -;::-'"
~
III
~~~ 0.. .. .. -
* I--t---+-I--Ir-_f- -:==::: =: "1":1: It,
'0
0.' - 1 i -1----1- - ------ -- ---.' '::., .- .
a::
c:
Q)
E
.,
Q)
E 0.04 1--+-1-+-+-' --- - - - - ---. - - - iJ - ..
III I--I--I--t-I-~'- --
0_02 :- - t --t---II--H- ' - - -- --.--- --- -. -_._ -

0.01 L.....L......J.._...L-...L.-'-_ --'_..J - ' - --'-----'---'-. - -


2 '0 20 40 100 200 400

Ti me (m;ns)

Fig. 14 Creep behaviour in side friction pile


are being dominated by the upper and most recent layers of fill, a feature that could profitably
be examined funher (noting, however, in some other cases that creep seems to have been
initiated during filling).
The principles applied here to dams apply equally to creep wherever it may occur, as in
slopes, foundations, etc. In all these situations, creep differs from the more familiar
mechanisms of settlement in being irregular and subject to discontinuities at any time. Unless
these discontinuities can be identified and isolated, there is no procedure that can be trusted to
prediet creep movement, but the isolation of just one discontinuity allows predictions to be
made with very substantially increased accuracy.
236

Al'know)edgement

The analysis of dam settlements herein has been made possible by the kindness of the Hydro-
Electric Commission of Tasmania, in providing access to their records. Permission to publish
is gratefully acknowledged, but interpretations thereon are those of the Writer. In addition,
acknowledgement is made to the Rural Water Commission of Victoria for permission to publish
Fig. I and to the Melbourne and Metropolitan Board of Works for permission to publish Figs.
3 and 5. Appreciation is also expressed to Dr. A F. Williams for access to data from his
doctoral investigations.

References

Bishop, A W. (1974). The strength of crustal materials. Engineering Geology (Special Issue),
8:139-153.
Bishop, AW. and Lovenbury, H.T. (1969). Creep characteristics of two undisturbed clays.
Proceedings, 7th lnt. Conf. Soil Mech. and Found. Engg. (Mexico), 1:29-37.
Cole, B.A. (1971). Wilmot rocUill dam - Concrete face deflections. ANCOLD Bulletin, No.
33, 19-26.
Cole, B.A and Fone, P.J.E. (1979). Repair of Scotts Peak Dam, Tasmania. 13th Congr. Int.
Comm. Large Dams (New Delhi), 2: 211-23l.
Fitzpatrick, M.D., Liggins, T.B., Lack, L.S. and Knoop, B.P. (1973). Instrumentation and
performance of Cethana Dam. 11th Congr. Int. Comm. Large Dams (Madrid), 3: 145-164.
Growdon, J.P. (1958). Rockfill dams with sloping clay cores. Proceedings, ASCE, v. 84, no.
P04, Papcr 1743. See also: Rockfill dams: Performance of seven sloping core dams, Papcr
1744.
Hydro-Electric Commission of Tasmania (1975). Gordon River Power Developmcnt (Part II).
ANCOLD Bulletin, No. 41, 22-26.
Meigh, A.C. (1976). The triassic rocks, with particular reference to predicted and observed
performance of somc major foundations. Sixteenth Rankine Lecture. Geotechnique, 26:3,
pp 391-452.
Mesri, G. and Godlewski, P.M. (1977). Time-and Stress-Compressibility Interrelationship.
Procccdings, ASCE, v. 103, no. GT5, pp. 417-430 (Paper 12910).
Mitchell, W.R., Fidler, J. and Fitzpatrick, M.D. (1968). Rowallan and Parangana rockfill dams.
J. 111st. Engrs. Australia, 40: 239-249.
Mitchell, J.K., Campanella, R.G. and Singh, A (1968). Soil creep as a rate process.
Proceedings, ASCE v. 94 no. SM1, 231-253, Paper 575l.
Mitchell, W.R. alJd Fitzpatrick, M.D. (1979). An incident at Rowallan Dam. 13th Congr lnt.
Comm. Large Dams (New Delhi) 2: 195-210.
Parkin, AK. (1971). Application of rate analysis to some settlement problems involving creep.
Proceedings. 1st Australia-New Zealand Conf. on Geomechanics (Melbourne), I: 138-143.
Parkin, A.K. (1977). The compression of rockfill. Australian Geomech. J., G7 pp 33-39.
Parkin, A.K. (1978). Coefficient of consolidation by the velocity method. Geotechnique 28:4:
472-474.
Parkin, AK. (1981). Consolidation analysis by the velocity method. Proceedings, lOth Inl.
COI1f. Soil Mcch. and Found. Engg. (Stockholm), 4 : 723-726.
237

Parkin, A.K. (1983). Strength and compressibility of rocklill. Blue Rock Dam Project, Tanjil
River. RepOlt to State Rivers and Water Supply Commission of Victoria, No. 83/1, Dept. of
Civil Eng., Monash University.
Parkin, A.K. (1985). Settlement rate behaviour of some fill dams in Australia. Proceedings,
11th Inl. Conf. Soil Mech. and Found. Engg. (San Francisco), 4 : 2007-2010.
Pigeon, Y. (1969). The compressibility of rocklill. Ph.D. Thesis, University of London.
Poulos, H.G., de Ambrosis, L.P. and Davis, E.H. (1976). Method of calculating long-tenn
creep settlements. Proceedings, ASCE, v. 102, no. GT7 pp 787-804, Paper 12273.
Pusch, R. and Fe1tham, P. (1980). A stochastic model for the creep of soils. Geotechnique,
30:4, pp 497-506.
Sowers, G.F., Williams, R.C. and Wallace, T.S. (1965). Compressibility of broken rock and
the settlement of rockfills. Proceedings, 6th Int. Conf. Soil Mech. and Found. Engg., (Paris),
1:335-340.
Wahls, H.E. (1962). Analysis of primary and secondary consolidation. Proceedings, ASCE,
vol. 88 no. SM6, pp 207-231, Paper 3373.
Williams, A.F. (1980). The design and performance of piles socketed into weak rock. Ph.D.
Thesis, Monash University, Australia
See also: Williams, A.F. (1980). Proceedings, 3rd Australia - New Zealand Conf. on
Geomechanics (Wellington), 1:87-94.
See also: Williams, A.F., Johnston, LW. and Donald, LB. (1980). Proceedings, Int. Conf.
Struct. Foundations on Rock (Sydney), 1: 327-347.
CH APTE R 10
FILTERS AND DRAINS
j . BRAUNS

1. I n tro duc ti o n

Recent l y, ICOLD publ ished a review of " Conventional Methods


in Dam Con struction " (ICOLD 1990). In the chapter devoted to
filters a nd drains we find t h e following formulations :
" ... Granular filters are required to prevent the smaller
size fractions from migrating from adjacent material ...
The Terzaghi criteria for grain sizes is generally followed,
being:
JIll
4 and n8SF < 4 .. "
S
It is not surprising that there is a great deal of
confusion in the understanding of filter problems, if this
matter is dealt wit h in such a way as in the ICOLD article .
First of all, one should not use the words "filter" and
" drain " in one breath, as we too often do. As is clearly to
be seen from figure 1, filter and drain functions are in no
way required at the same time in actual si tuations. In the
first case shown, a drain i s required, having a sufficient
drainage capacity to drain off the seepage quantity of
percolating water. The material of the drain, in fact, has
to meet the filter requirements so as to prevent erosion in
the d a rn . In the second case. a filter is required in order
to prevent contact erosion between the darn material and the
toe drain . By no means is a drainage capacity of this filter
zone required here. The filter and drainage functions should
clearly be kept apart from each other and the actual
requirements i n each individual case have to be properly
checked.
It is therefore not adequate to refer to " Terzaghi' s
filter criteria " as quoted above, since the first criterion
give n refers to the drainage function whereas only the
second criterion is the one referring to the filter stabi-
lity problem.
2"
E. Ma fanho das N~,'ts (~d.J. Ad"ancn in RIXifrll Slr"c"'r~s. 239-268.
C 1991 Klu ...tr Academic Publishers.
240

~ ..::.

~77
DRAIN

HOMOGENEOUS DAM WITH DRAIN

?// 777

HOMOGENEOUS DAM WITH TOE DRAIN AND FILTER

Figure 1. Illustration of drain and filter functions in


dams

Not even the introductory remark quoted is correct when


mentioning the "smaller size fractions from migrating from
adjacent material". As is well known, one has to rely on the
internal filter stability (or "autostability") of the finer
material in a base/filter combination of two soils in
contact. Thus, it is sufficient if the filter is able to
retain the coarse fractions of the base material. This can
be seen directly from Terzaghi's filter criterion D15 F/D85 s
< 4 (see above) in that it uses D85 S (S denoting the base
material) to characterise the base material, that is the
grain size of the 85 %-fractile of this soil.
In the following presentation of the present practice of
filter design, the index F is used to describe the kilter
material, the index B is used to denote the Qase material,
as is common in current literature relevant to this theme.
241

2. Present Practice of Filter Design

2.1 GEOMETRICAL CRITERIA


One of the first attempts to define conditions of filter-
stability seem to have been made by a man named Prinz in
1923 in connection with filters around wells. He figured out
a critical relation of grain diameters of filter and base
material on the basis of geometrical aspects regarding
spherical grains. A number of authors have treated the
problem in this way; others performed experiments using more
or less uniform soils in column tests.
In figure 2, a selected number of geometrical filter
criteria is listed up as they were developed since the
twenties.

mass fracti Ie
05 15 50 85 100

PRINZ (1923) 4.42

BERTRAM (1940) ~~~ ___ ~ni!,. B

TERZAGHI/PECK (1948)

2.4 (l(X)se)
SICHARDT (1952) 6.5 (dense)

US CORPS OF ~-- _ 5 (safe) -


ENGINEERS (1953) 4~ ~ 25

LEATHERWOOD/ ~---~ 4.1 --- ----


PETERSON (1954) 4~ c 5.3

KARPOFF (1955) 12 - 40
o " 5 - 10 (unif F)
'we1 grade:
(USBR) o - 12 - 58 I B, F
ZWECK/DAVIDENKOFF
4 .: 5 - 10 (unif B)
(1956)

ZWECK (1959)

SHERARD ET AI.. (1984) -< -, 9 -- - ---------<

~ "4 -------i lunif. B


KENNEY ET AL. (1985)
~ . 5 --~ well graded F

FIL T E R . BASE
F~-- sIze f a c t o r - ,

Figure 2. Graphical presentation of various geometrical


filter criteria
242

A direct quantitative comparison of the criteria is not


possible, for a number of reasons:
- the failure conditions in tests have been defined by the
authors in different ways,
- different points of the gradation curves (d15, d50, d85)
have been chosen for the representative grain diameter,
- safety factors have been introduced by some of the authors
when defining their ratio of grain sizes; other authors
defined the ratio under limit state conditions,
- some authors performed experiments, others made theoreti-
cal analyses, as has been mentioned already.
As far as coarse grained soils are regarded, systematic
tests have recently been performed by Sherard et ale (1984).
These authors ascertained from their tests, that a "washing
through" of the base soil through the filter occurs, if
d15 F/d85 B reaches 9. This corresponds well with the tests
made by Bertram in 1940 (see figure 2), on which the
criterion published by Therzaghi and Peck in 1948 may also
have been based, which is now commonly called the Terzaghi
filter criterion. This criterion thus includes a certain
safety factor.
witt (1986) has shown that the experimental investigations
of Vaughan & Soares (1982) on the filter stability of silty
materials can also be expressed in terms of d15 F /d85 B . The
results are shown in figure 3, together with tfie critical
ratio of 9 after Sherard et ale or Bertram.

SIL T SAND
medium coarse fine me<lium coor~
III
20 I

'"'"
"0

;,- 15
"'<'VAUGHAN and
SOARES
EROSIONS--
"0 ~
~, I
10 ... /,SHERARD etal
r-- --- - - - - ... ... ...
t---. -.

--
-.
-.
5 ~ FILTRATION

1 I
0,006 0.D2 0,06 0,2 0,6 2
d858lmml
OF BASE MATERIAL

Figure 3. Results of filter tests of Vaughan/Soares and


Sherard et ale (after witt 1986)
As can be seen in figure 3, the critical ratio of grain
sizes increases beyond 10, as the base material becomes
increasingly finer. This means, that the problem of filter
243

stability for silts and clays is no longer a purely


geometrical problem. In other words: fine-grained base
materials prove to be filterstable under a high hydraulic
load, even if a washing through of the base material through
the filter material is geometrically possible.
In most of the criteria mentioned here, the soils adjacent
to one another are represented by a distinct grain size d15
or d85 as in the case of the Terzaghi criterion. The major
portion of the geometrical criteria available from litera-
ture use d50 as the average value to characterize a granular
soil. As long as the soils in question are uniform, the size
of the fractile diameter chosen is of no particular impor-
tance. But when well graded soils are investigated, a second
parameter such as the coefficient of uniformity Cu comes
into play, if d50 is used as a significant fractile
diameter.

m 60
C)

'" I I
-
EROSION C U' B=10), ~7/-'/
"0
50 /"/;;-- ,,""2
"-
C)
..-:;.--
..-:::::-<1;~7____
7
--(1
'"
y -::::::~ ___
"0
40 ZlEM5 / __ --
.J;.~ __

; / --
/ / .........
;//,;--::y"/
-------'//
30
;/
I
1:%
FILTRATIONI
20
yY" 7 THANIKACHALAM and 5AKTHIVAOIVEL.
// d50 F Id50 a =2,4'C U,F-8
1O

OL-~~--~--~--~--~~~~--~---
1 2 6 10 12 14 16 18 20
COEFF OF UNIFORMITY OF FILTER CU,F

Figure 4. criteria after Ziems (1968) and Thanikacha-


lam/Sakthivadivel (1974) for well graded soils, as compared
by witt (1986)

In figure 4, this is shown with examples after Ziems (1968)


and Thanikachalam/Sakthivadivel (1974). The critical ratio
d50 E/d50 B is significantly dependent on the coefficients of
uniformi~y of both the filter and the base materials.
When the results of these authors and others are studied
in detail, one comes to the conclusion that the fractile
diameter d50 is not a very good means to characterise a soil
for a filter criterion. wittmann (1980) has shown by
theoretical analysis, that the diameter of a high fractile
is the best means to characterise the base material as long
as the soil is autostable. This can easily be made evident,
if we imagine the flow of a graded soil through a mesh or
sieve: As soon as some coarse particles have blocked the
sieve fines can no longer escape through it.
244

There is a further question on the other hand, regarding the


best fractile diameter to reasonably characterise the pore
system of the filter material, through which the base soil
may try to escape. As one may deduct from Hazen's law, which
uses dID as a significant value to characterise
permeability, the diameter of a low fractile is a good means
to quantify the pore size of a granular mass. In fact, d15
is used in Terzaghis filter criterion.
As regards the first aspect - d85 as a significant value
for the base soil - there is one important exception:

SAND GRAVEL
~----,-----,-----~----'------r----~~OB
m c m c
1- 100
I
<.:>
w 85
3: 80
>-
m
0::
w 60
z
LL i
I I
";f.
1.0

20
15

GRAIN SIZE d [mml

Figure 5. The use of the Terzaghi filter criterion for two-


component base soil
In certain cases, the use of two-component soils cannot be
avoided in embankment dam construction. This happens, for
instance, if slope debris is to be used, which contains a
certain amount of rock fragments. In other cases, desinte-
grated rock masses result in fills with a gap-graded
gradation. To complete the discussion here, it should be
mentioned (because sometimes it has been overlooked), that
- before application of the filter criterion - the gradation
of the fine matrix has to be determined as is shown in
figure 5. This is necessary because two-component soils of
this sort are far from being internally filter-stable
(autostable) .
Unfortunately, even some reasonably well graded materials
(without a "gap" in the gradation) prove to be internally
unstable and can show what is called "suffosion", as
investigations by Russian researchers have shown (Lubockov
1969). In other words: The fines of such materials can be
245

washed out of the graded material when percolated with


water.

85

'+--+l---t-1IR DISc
DasF

g,Lo,--~~----~~~~~---'I~,O----------~10

Figure 6. Method for analysing the internal stability of


soils by splitting the gradation (after Sherard 1981)

Sherard (1979/1981) has made a proposal to check autostabi-


lity of well graded soils by splitting the gradation into
components at various points and comparing d85 f . with
d15 with the rule I = d15 /d85 f ~ 5, 15r all
splI~~IR~ points (see figu~e 6 fo~ an example). Other
authors have worked further on this problem and discovered,
that autostability of a graded soil is given, if the ratios
d85/d50, d50/d35 and d35/d15 are all ~ 5.
Generally, it can be said, that "the wider the grading and
- thus - the larger the coefficient of uniformity (d60/d10),
the more likely it is that the soil will be internally
unstable" (see ISSMFE-Report on Filters).
If one is faced with a situation, where a filter has to be
designed to safely protect a soil, which is not autostable,
it should be made to suit the fine fraction particles, which
are expected to be moveable, as shown in figure 6 for a gap-
graded soil.
As regards the autostability of filter materials them-
selves, one should make sure that filter gradations meet the
requirements for autostability (see above). In general, the
maximum grain size of graded filters should be limited to
approximately 50 (or max 100) mm. This helps to reduce the
tendency of segregation during placement and - thus to
ensure a sufficient degree of homogeneity of a filter zone.
246

2.2 SCATTERING OF GRADATIONS

o em 10 20 30 40 50 60

Figure 7. Surface of a test embankment showing local


inhomogeneities

The geometrical criteria described above assume, that a soil


or a filter material can be defined with a unique gradation
curve. But, in actual cases the gradation of soil masses
scatters. This - in general - applies to fine grained soils
(base materials in a base/filter combination), for filter
materials (to a limited extent, at least), and - particu-
larly - for rockfills. This is important because the problem
of filter stability or filter instability occurs in various
interfaces between darn zones, not only at the surface of a
"filter" in the strict sense.
Safe application of the geometrical filter rules requires,
that filter stability is given in all possible combinations
of the various gradations within the gradation bands of (i)
the soil and (ii) the filter (see figure 8 for this).
This means that the filter rule must be applied to the
left borderline of the soil (base) and to the right
borderline of the filter gradation.
247

SAND GRAVEL COB


I

tr--
~
0 f m c f m c
w 100 1
~ ~>"
>- 85 --+-- -
co 80 BAND OF - k~ ~ 1-- l- f--
cr
w
z
LL
OFBA'j I~f\~-
GRADATIONS~

60 SOIL v(")
I;j
11
::) :..- BORDER - LI NE
FOR
l
ALLOWABLE FILTER f--
;;e L\'" GRADATIONS I--
~
~O ~ t-- r'
~ }
I
-
V
20 -
15 -- I-
I
- /--< l- f- - -- . -'- --
1 II
----, '--

o ~
006 01 0.2
I 0.6 to I 2 6 10 20 60 100
RELEVANT d85 S d15 F
GRAIN SIZE d [mm]

Figure 8. Terzaghi's criterion of filter stability applied


to a band of base soil gradations
2.3 FINEST FRACTION IN FILTER MATERIALS

As is well known, the fine fractions of a soil define its


pore sizes which in turn control its filter capacity against
a finer base material. Thus, when applying a filter crite-
rion to a gradation of a filter material, one must rely on
the fact that the fines in the filter do not move out. The
question of autostabilty becomes very important here: If the
gradation of a filter has a fine tail which proves to be
internally unstable, these fines should either be removed
from the filter material or - at least be ignored
(analytically) in the design process. Sherard's method men-
tioned above may be regarded as a useful tool to determine
the unstable fraction.
2.4 FILTERS FOR COHESIVE SOILS

Truly cohesive soils which are "not dispersive" prove to be


stable against erosion into a filter even if the geometrical
criteria are not met. This has been mentioned already in
connection with the test results of Vaughan/Soares on silts
and clayey silts. As a matter of fact, the risk of erosion
or piping in a base/filter contact with a cohesive base
material is mainly controlled by (i) the degree of non-
dispersivity (which is difficult to define) and (ii) the
hydrodynamic load (that is: hydraulic gradient or flow
velocity).
248

A safe filter design to safely protect very fine cohesive


soils calls for mUlti-stage filters with fine sand in the
first position. In this connection it is useful to learn
from the investigations of Vaughan/Soares that very fine
sands are able to protect even dispersive clayey soils under
ratios d15 F/d85 R as much as 10, because the fine materials
(base) appear in a flocculated form if suspended in water or
after erosion in an erosion channel.

2.5 HYDRAULIC CRITERIA

Hydraulic criteria - in connection with erosion in inter-


faces between different soils - may be involved in situat-
ions, where the meeting of geometrical filter criteria (say:
by technical means) is not in our hands. A situation of this
kind may occur, for instance, if geometrically unstable

Figure 9. stratified sedimentary granular soil (sediments of


Rhine River)
249

layers of soil are found in the ground (compare figure 9)


and percolation is induced by some construction activity,
say through construction of a dam. The phenomena and
quantitative relations, which govern such situations, are
not of interest in connection with design rules. But, it is
interesting to look at these phenomena - at least in a
simplified and basic form -, in order to obtain a feeling
for possible related risks.
In fact, there is a broad variety of situations, which may
be encountered, as regards the orientation of flow, on the
one hand, and of the base/filter interface, on the other.
Let us take cases 2a and 3a of figure 10 only in order to
demonstrate a few effects.

~VB

dVF
10 1b 20 2b 30 3b

Figure 10. variations in the orientation of flow and of the


base/filter interface
In case 2a (figure 10), the seepage flow is directed upward,
from the base soil into the coarser one, say a "filter". The
mean velocities in both soils are the same in this case.
Under such conditions the hydraulic gradients are inversely
proportional to the coefficients of permeability, which
means that the base soil is under a high hydrodynamic load;
this load is directed upward, that is against the direction
of the buoyancy weight of the soil.
For the case described, a hydraulic filter criterion in
terms of grain sizes has been derived by Ziems (1968). This
criterion is given in figure 11, which also shows the
respective formula. As is to be expected, the critical
gradient is a function of the grain size of the underlying
base material (d10 is used by Ziems), and of a grain size
relation between ~he two soils in contact (A50 = d50 F/d50 B
is used by Ziems for this) .
250

i- \\
I CRIT

:
I-
14
~ 1\ I
I
jII CRIT= 0, 66 +
I-
z
lJj 12 6
I d10 2 A50 2
0
<
\

\J
0:: 10 ---- --I
\
lC>

t
U
mSu
::; 6 ------1I
=>
<
0::--'
0<
I
I
g Su \ I,
>-- 6

i\
--
:r:ffi fS \ , , * ' O S = 0,02 mm
I-
--'< 4 1\
<::E -.....
u
-lJj Il~10s=O,~m~1Os =O,06mm I'-..
~~l~dlOs.:,E,6mm l-
~Ul I--
0::< 2 I
urn
0 I I j
0 4 [10 20 30 40 50 60 70 60 90 10 o
d10S=2mm A 50 = d 50 f
d SOB

Figure 11. Hydraulic criterion for flow across an unstable


base/filter interface (after Ziems 1968)
For very large A50-values, I . takes a value somewhat
below unity, because then the pf56tem degenerates to that of
hydraulic heave in the finer layer. On the other hand, for
A50 towards 4, geometrically filterstable conditions are
attained, where the soil combination is stable under very
high (or even infinite) hydraulic gradients. Moreover, the
grain size of the base material is the governing factor.
Using dl0 B as a significant grain size to characterise the
base soil is not in contradiction to what has been said in
connection with hydraulic criteria under non-stable condi-
tions here, because this is not a geometrical problem.
Let us now consider case 3a of figure 10 for a moment,
where the seepage flow is horizontal and parallel to the
interface between the base and "filter" soils. Under such
conditions, the hydraulic gradient is the same in both
materials, but there is a significant difference in the flow
velocities, which are proportional to the coefficient of
permeability. We have carried out some tests in Karlsruhe to
investigate such situations (Brauns 1985). From the results
and their analysation, we were able to derive a criterion
for a critical hydraulic gradient, which causes erosion of
the fine base material along the interface between the two
soils in contact.
The critical gradient is shown in figure 12 in the
function of the mean diameter d50 B of the base material and
of the hydraulically relevant grain diameter d WF of the
"filter".
251

with i 1 ... n fractions of a gradation


tip , mass percentage of fraction i
d,mJ. mean diameter of fraction i
J.
Tests and research were carried out for uniform soils only,
so that we do not have results for graded materials, so far.

iF erot dwF = 5 mm
0.6 r---;========;-------:T--,
Frc~t =0.65 =const.
l.e = 26.5 kN 1m 3
0.5 nF = O. 39

ICrnd.: nFd15F I d85e ~ 10 I


0.4 7.5

0.3
10

0.2 12.5
15
17.5
20
0.1 25

o~~~~=-~~~~
o 0.1 0.2 0.3 0.4 0.5 0.6 0.7
[mml d50s
Figure 12. Hydraulic criterion for flow along an unstable
base/filter interface (after Brauns 1985)
As can be seen in figure 12, very low hydraulic gradients
are critical, if the "filter" is coarse and the base
material is fine. This can easily be understood, if we
regard the high flow velocities in the pore system of coarse
materials on the one hand, and the low critical drag forces
required to move fine particles (base soil), on the other.
Please note, that in both figures, 11 and 12, critical
gradients are presented, which do not include any safety
factor as do the geometric filter criteria discussed
earlier.
At the end of this section concerning the present practice
of filter design, it should be emphasized that the diagrams
in figure 11 und 12 are only given here to draw the readers
attention to quantitative relations, which govern the risk
of erosion, if percolation of unstable soil combinations
cannot be prevented. These diagrams should not be understood
as design diagrams.
252

3. Recent investigations on the problem of filter


stability

In this section, we deal with recent investigations on


filter design problems. We will restrict ourselves to the
problems of filter stability between two adjacent soils. The
question of internal stability of the two soils in question
will not be considered, in other words: we assume autostabi-
lity to be given.
In looking in more detail into the conditions of
available grain sizes (base material) and pore sizes (filter
material) in the region of contact between two soils, it
has been recognized that matters are a little complex and
call for a statistical approach. In fact, the local grain
size distribution of the base material and the pore size
distribution of the filter are of fundamental importance in
what we call the stability against contact erosion in a soil
interface.
DOODG
GRAIN SIZE OF o QQ
BASE 000000
FILTER = SCREEN
g~cPoO
WITH VARIABLE
PORE SIZES PORE SIZES
MESH SIZE ~_TE_R _____ OF FILTER

PROBABILITY OF PROBABILITY OF
FAILURE Pt = 100". FAILURE Pt= 0

Figure 13.: Opposing conditions of failure probability of a


base/filter combination
In the ISSMFE-Report on Filters, one can find the following
very clear explanations (compare fig. 13):
"Where the largest grains of the base are smaller than the
smallest pore of the filter the probability of failure is
100 %. Conversely where the smallest grains of the base
material are larger than the voids of the filter the
probability of failure will be zero. Since the latter can
only be possible where the filter is either finer or
identical to the base material, and by definition the filter
must be coarser than the base, the probability of migration
of particles from the base soil is therefore of finite
magnitUde for any filter system. Selection criteria are
therefore designed to limit migration to acceptable limits."
This is a very clear definition of the problem and of its
statistical nature.
253

;f!. 100 DISTRIBUTION """",-


:_ 80 BY NUMBER-:o'Y'--_ _ _-/"'-- L _ _-----j

o /1
- 60 -+---- ---,(''---
E,
a~ 40 + ___
I
----/-c--+_ _~--_-----j
I
2Df-----t---c,,?-;---+----+---

m
dID d30 d60 d90 d-
?6~.

~,~q ~'
.::1_

"0 ..
.. '

c;~~~;:: ::.~~,. ;~
:~.:?,:;, 'e S?
SCREEN WITH d 30 d 60 d 90 LOST MASS
MESH SIZE
EQUAL TO '.:.~.o.: OF BASE MAT'ERIAL
~. :... ~~ ?~j~): "_9,

Figure 14. Base losses to block a filter with a defined pore


size (after witt 1986)
Let us in a first step - assume that a filter may be
represented by a screen with a defined pore size. Figure 14
shows four selected cases, where the pore size of the filter
screen is equal to dl0, d30, d60 and d90 of the base
material, respectively. The quantity of base material pas-
sing through the filter until it is blocked is variable
within wide borders, depending on the relative fractile
diameter of the screen size.
The pore system of a true filter cannot simply be represen-
ted by a screen of defined mesh size. The pore sizes, and in
particular the so called pore constrictions scatter statis-
tically. Furthermore, along a flow path in a filter, there
is a sequence of pores and of pore constrictions with a
statistical distribution of sizes. Silveira (1965) may have
been the first, to point to this fact. These statistical
properties of the pore system of a filter are involved here,
since we have seen that 100 % safety - in other words: zero-
probability of local failure - in a base/filter interface
does not exist.
Figure 15 gives an example taken from the work by witt
(1986) who has made measurements on actual models of pore
skeletons. Tne distribution diagram of the figure shows
- the grain size distribution of a filter 8 to 63 mm by mass
on the right side,
- then the distribution of grain sizes by number,
- on the third line from the right: the distribution of the
maximum sizes of constrictions of the pores in the entire
mass
- and finally: a sequence of curves representing the sizes
of largest pore contrictions in a function of the length
of the flow path or of the pore canal. NE denotes the
number of pore constrictions along the flow path.
254

DISTRIBUTION OF MAX
PORE CONSTRICTIONS
I N, ~ NUMBER OF CONSTRICTIONS
ALONG FLOW PATH I
A- ___ ~ __ ~ __ ,
1.0
0

85%
Ul
UJ 0.8
'"
;".
z
UJ
u 0.6
a:
UJ
"-
LL
0
0.4
::E
:::>
Ul

31. 5 63

40 6Ci 80 100
0.155 do dl00
DIAME TER d [mm I

Figure 15. Cumulative distributions of grain sizes of a


filter material (by mass and by number) and of pore
constriction sizes, with number of constrictions (along flow
path) as parameter (after witt 1986)
Two things can be concluded from this material:
a) the significant size of pore constrictions is a
function of the thickness of a filter,
b) the distribution of the significant pore sizes is
relatively weak; in other words: the significant pore
sizes are relatively uniform after a certain penetra-
tion depth into a filter.
The latter point in turn means that a filter of a certain
width has a well defined significant pore size and will
completely retain a uniform granular mass with a particle
size larger than this pore size, but will let pass comple-
tely a mass with a particle size smaller than this pore
size.
Let us, on the other hand, regard the thin dotted line
drawn in figure 15. It shows a soil which is filterstable
against the filter according to Terzaghi's criterion. If we
compare this line with the lines representing the filter
pore system for different numbers of NE that is for
different thicknesses of the filter - we see, that a good
overlap of the pore size line and the base gradation line is
only given for a certain number of NE , that is for a certain
thickness of the filter. If the graaation of the base were
on the left side of the pore size distribution, the
condition for the complete washing through of the base
material would then be given. On the other hand, the
255

relative position of the base gradation in regard to the


different pore size distribution lines defines the depth to
which base material can migrate into a filter.
The theoretical and experimental investigations performed
on this problem have shown that - with increasing thick-
ness - the significant pore sizes of a certain filter
material tend towards a certain value, which we called the
"effective filter size (d *)" of a filter size (Brauns &
witt 1987). As could be shoen this effective filter size can
be calculated from d50 of the grain size distribution by
number of grains:
average grain size by number)

BASE FILTER

ORIGINAL
INTERFACE
~~

ry
Goa

Figure 16. Formation of graded filter at interface (after


ISSMFE-Report on Filters)

From all this, we see now that formation of a graded filter


at the interface of two soils in contact is inevitable and
that it is true what is said in the ISSMFE-Report on
Filters:" selection criteria are therefore designed to
limit migration to acceptable limits ... "
At this stage, the question comes up as to what might be
the criteria of "acceptability" of unavoidable particle
migration.
256

80r------------------------"

60 lesl s analysIs

.
l 0~~~~B~S====~gO====~9~5----~IOO
QB "' {d~ I I % I

r
1.... .~ ~~. B!!E .. :' . ":.". INm.
bQse moss losl
penelralian L . : .. : ... +-;-:;+-;-;:'" FACE
depth If
.. ..=....-.:
..-.~ ...
. ...-.-:-.. . ..
. .:.-.-.
1000,------------------------;.--,

anolysls
~I~ BOO Irsls
600

4 00
3 JO
200
. '.
- - - - - - - - - - - - - - - - -. -

85 90 95 100
Ce,m {d~ J I'!,}

Figure 17, Base mass loss and depth of penetration into a


filter in a function of the effective filter size as
compared to gradation of base soil (after witt 1986)

As far as we can see, there are - all in all - three cri-


teria to look for:
1. The filter must be of a gradation and thickness so that
no continuous migration of base particles can occur
(this could be called "stability criterion") ,
2. Provided that criterion No. 1 is fulfilled; the extent
of base material losses must be tolerable with respect
to the purpose of the zone made of this material
(example: the volumetric losses of base material must
not lead to piping in a sealing zone of a dam, say a
thin core of fine grained material; this could be
called "piping criterion").
3. Provided that criteria Nos. 1 and 2 are fulfilled; the
extent of infiltration of base material into the filter
must be tolerable with respect to the purpose of the
filter zone (example: the depth of infiltration into
257

the filter must not be such that the reduction of the


permeability of the filter leads to misbehaviour of
this zone; this could be called "drainage criterion",
but is to be seen in close relation to the filter
problem)
criterion No.1, the "stability criterion", is an absolute
must; there is no way around. criteria Nos. 2 and 3 are
related to each other, as is shown in figure 17.
It it easy to see from the sketch in the centre of the
figure that these criteria are closely linked to one
another: The mass mB lost creates a cavity on the base side;
the infiltration of base material into the filter zone
creates a zone of reduced permeability, the depth of which
is I (this depth must be smaller - at least - than the
thickRess of the filter because of criterion No.1).
The upper diagram deals with the base material losses: It
shows the results of theoretical and experimental investiga-
tions in terms of a relatio~ between lost mass of base soil
and relative position of d (of filter) as compared to the
gradation of the base mateFial (compare figure 18 for this).

SUM OF PERCENTAGES Q
10~------------~----------~'--T--,

08

06
05

04

02

31 5 63

1 2 I 4 6 8 20 40 60 80 100
EFFECTIVE FILTER SIZE: dt:r;:n dk
DIAMETER d [mml

Figure 18. Definition of QB,m (d p * ) for a base/filter com-


bination
As is to be expected, th~ extent of base mass losses
increases with increasing d of the filter (figure 17). In
particular, these losses Pincrease dramatically, if the
effective filter diameter exceeds d95 of the base material.
Under these circumstances, almost a complete wash through of
the base material is about to occur.
258

In accordance, the penetration depth 1 of fines into the


filter increases dramatically as the eff~ctive filter dia-
meter approaches d95 of the base material (lower diagram in
figure 17).
Both effects - base losses and.penetration into filter
can be limited by restricting d to say d97 % of the base
material, in other words: the di~tance between the gradation
of the filter and that of the base is restricted to a
certain value, as is done in Terzaghi's law. A restriction
law of this type can be formulated from the relations shown
till now.
Up to here, we have described a base material with a
unique gradation curve. We have also described a filter with
a unique gradation curve.

50 40
If) d 85:
z $.
0 IN =1281 !.I(dIS} =57.94"",
t= 40
30 t; (deS) = 9.37mm
~ V(d8S} = 16.2'.
Z
II:: W
l:ll 30 :::l
d 15
In
0 20 ~
II::
~ 20 U.

II::
W 10
~ 10
:::l
Z
0
0.5 2 4 8 16 315 63 100
d15,d85 [mml

Figure 19. Histograms of a rockfill mass as derived from


numerous samples taken from a test fill

However, as we all know and as has already been shown (see


figure 7), we are normally faced with a certain amount of
scatter in the gradation of granular materials. Figure 19
gives - as an example - the histograms of d15 and d85 of a
large number of samples taken from a test fill. Both d15 and
d85 proved to be normally distributed, in the case studied
in detail here. This shows that the problem of filter
stability is a statistical problem not only with regard to
the process of filtration along individual flow paths, but
also with respect to the scattering of soil properties from
place to place, for example in a large contact area
(interface) between two adjacent zones of a dam.
In using the concept of limited base losses and limited
material penetration into the filter - on the one hand ,
and in applying some statistics in regard to the scattering
in the local gradation of the materials in question - on the
other, the filter stability of two materials in contact can
be described quantitatively in terms of the probablitity of
259

failure. It is not possible to explain this in detail here,


due to lack of space. However, it is important for an
engineer faced with actual problems that filter criteria can
be derived from such an analysis, which are still relatively
easy to handle and which make it possible to quantify the
risk or probability of failure of the system regarded.
Reasonable filter criteria of this ~~rt, which incorporate
a probability of failure of p = 10 , are presented in
figure 20. The probabilistic an~lysis showed that one should
distinguish between "fine grained filters" (with d5 F < 0,5
mm) and "coarse grained filters" (with d5 ~ 0,5 mm). For
the fine grained filters, d95 B of the bas~ material should
be taken as the significant grain size, d85 should be chosen
for coarser filter material.
The analysis further revealed, that a critical ratio
dx jdy ~ 2.5 should generally be observed. The relevant
frKcti~es of the filter are a function of the coefficient of
uniformity of the material (for details see figure 20).
Please note that a minimum filter thickness of I
330 d5 F has been introduced into the theoretical analysation
of the problem. This calls for instance, for a filter
thickness of 2 m for a material with d5 F of 6 mm. Thus, when
dealing with filters, we should always relate the filter
thickness to its grain size and, in general, be careful with
thin filters.
It should be mentioned here, at the end of this chapter,
that the probabilistic theory developed by witt (1986),
which leads to the filter criteria mentioned above, is only
one of various possible approaches to the problem. On the
other hand, such a theory makes matters more rational, but
not necessarily easier to deal with. But there is no way
around statistics in problems of this type, unless we prefer
to ignore their proper nature.
260

I FINE-GRAINED FILTER (d5,< 0.5mmJ I


SAND GRAVEL

100
~

>-
III

~ 50

-30

-10
5
o
0.06 0.2 0.6 2 6 20 60
GRAIN SIZE d [mm J

COEFF I CrENT OF UNIFORMITY OF F IL TR

CU , F
( 3 3 , ~ 6 CU ,F 6
CU. f

r '"e-9 r " , . d ' ' ' t e r


dS r 0.5 mm
m" d 30 F
-
( 1.5
m" d 10F
---
~ 1.5
m" d 5F
-
1.5
mln d 95 8 mln d 95 8 m, n d 9\ 8

COARSE - GRAINED FILTER (d5F~ 0 5mmJ I


SAND GRAVEL

100~~---+--~m~~--~r-~~--4-~~--h-~~
95
>-
i~
w
~
>-
tIl

50

-30

-10
5
O~ ____L-__ ~-L~~L- __ ~ ______L-__ ~

0.06 02 06 2 6 20 60
GRAIN SIZE d [mm J

COEFFICIENT OF UNifORMITY OF FILTER

I COiHse- g ralned fllter


d SF ?: 0 5 mm
max d 30F
::. 2.5
rna .. d 10F
- - - ( 1.\ 1. \

Figure 20. Filter criteria derived from probabilistic ana-


lysis (Witt 1986)
261

4. Drains
In the introduction to this contribution, we have tried to
separate properly the filter function from the drainage
function of design elements in dams. Up to now we have dealt
with the filtration aspect. This closing chapter will deal a
little with drainage layers and their drainage capacity.
For obvious reasons drainage layers are normally made of
pervious material, sometimes of clean gravel or even crushed
sound rock. In rockfill dams, the rockfill itself acts as
the drainage zone.

HOMOGENEOUS DAM.

ZONED DAM.

KOZENY-CASAGRANDE DUPUIT

hor=V2~+(L)2
kOr /<Or
:0 V2 ~
kDr
~ NEGLIGI BLE

Figure 21. Drainage of seepage water from dams


Figure 21 shows these two cases and also gives the well
known formula to determine the required height of the
draining zone. When designing a drain, one has to introduce
an additional factor of safety to compensate for the fact
that the coefficient of permeability of a soil is not a
parameter which is very precisely known. We should keep in
mind that the quantitative prediction of the geometry of a
seepage flow situation is - in general - of low reliability,
as soon as two or more zones of different permeability are
involved.
262

There is a large number of aspects which could be


discussed in connection with the design of drainage zones in
dams. Just to mention a few: effect of inclination of
impervious base, steep inclination of chimney drains, three
dimensional effects, provision of additional drain tubes to
improve capacity, vertical drainage wells in foundations.
In view of the limited space, we must refer to the
relevant literature for all these aspects. But with regard
to the main subject of this meeting, we should discuss
briefly the aspects of permeability of coarse grained
materials.

SAND GRAVEL
FINE MEDIUM COARSE FINE MEDIUM COARSE

"" 0, 2'~
~; CRIT
(d [mm] I w
w

u
---' 10-1
::J
<t
0::
o
VALID NOT VALID
~~
>-
I

10- 3 ~
0,06 1 0,2 3 50,6 1 2 3 5 6 20 3 560
EFFECTIVE GRAIN DIAMETER dw[mm]

Figure 22. Borderline between ranges of validity and non-


validity of Darcy's flow law
As we all know, Darcy's law of linear relationship between
filter velocity and hydraulic gradients is only valid up to
a certain Reynolds number. On the basis of some analyses
using pore canal models, one can determine that the critical
gradient, beyond which non-linear effects begin to play a
significant role, is proportional to the inverse of the
third power of the grain size d (see formula in figure 22).
This is a very strong relation ~hich can be illustrated by a
straight line in a double logarithmic diagram as shown. In
order to give an idea of the grain sizes along the abszissa
of the diagram, the corresponding ranges for sand and gravel
are shown at the top of the graph.
The borderline which separates the regions of validity
and non-validity of Darcy's law, is steep. As we can see,
for filter materials with d w in the fine and medium sand
263

region, hydraulic gradients of up to 10 and even more are


allowed without loss in apparent permeability due to non-
linear effects. At the other end, for drain materials with
d in the range of fine gravel, hydraulic gradients as low
a~ 1/100 or even 1/1000 are critical; beyond these
gradients the apparent permeability will be less that
expected according Darcy's law.
I/v [slem 1
0.06 --------

0.05
~3 d w 020. 52mm l
52
0.04 b 00014-
. cm 2
10 0 0.361

0.03

0.02 n =0,3B
x 0 0 0.36
+ 0 0 0.36
0.01

0 0 0.006 c~
0L--'~0.~5~-1~.0~-~1.~5--~2.~0-~2~.5~-"3.~0--~3.5
V [em Is 1

Figure 23. Non-linear law of flow through granular material


As is well known, the non-linear law of flow through
granular material can be expressed in a binominal form as
given in figure 23, which also shows results of tests as
compared with the straight line representing the result of
theoretical analysis. A curved line should even better fit
the experimental results, but a linear approximation as
supported by theory may also serve the needs in a first
step. The coefficents defining the binominal flow law are
"a" and "b". wittmann (1981) showed that there are relati-
vely distinct relations between the coefficients "a" and "b"
(which, by the way are not dimensionless) and the grain
size d .
Wit~mann collected as much data as possible from the
literature concerning results of permeability tests on
coarse grained materials and added his own results in
forming the diagram of figure 24. It shows the coefficients
"a" and "b" in a function of d. The lines are theoretical
results which correspond relatYvely well to the experimental
data, for which d ranges from 0.5 mm up to as much as
200 rom. w
264

ololm] bls'l m']


10 ' 1

10(. I

10'

10 -) 1 ,l.--,L------!--"--!---;------!---!----;-,-:-,-;--'7,---;, d..., !m J
101. ":0' 10
01 10C 1000d w imm]

Figure 24. Comparison of experimental with theoretical


values of coefficients of Forchheimer's law (Wittmann 1981)
It seems that the "permeability" (better: permeability
characteristics) of coarse granular materials can be fairly
well determined analytically from the given "a" and "b"
values.
As far as the drainage capacity of drainage layers is
concerned, it is not the non-linear flow law which we are
mostly interested in, but the reduction in apparent permea-
bility k). On the basis of the "a" and "b" values shown
above, we can derive a red~ction factor "K" to calculate
the apparent permeability (k ) on the basis of the one after
Darcy (k f ).
265

-
FACTOR OF REDUCTION K

1.0.
r--
K
I
ID.6mm

-- -
0.8

~\ ~ r--..... IK-~- 2 J dw=\DJ

--
- k - 1+v'1+C'I-d w3 '1

AD
0.6 f

0.1.
,\~ r--

JD-
-.............

0.2 ~
...........
r--
-........: ::::- r 10.0 r8D r6.D

0. 7 8 9 10.
0. 2 3 I. 5 6
(C'xD.2 mm- 3 ) HYDRAULIC GRADIENT

Figure 25. Factor of reduction in permeability due to non-


linear flow effects
This reduction factor depends, as is to be expected, (i) on
the hydraulic gradient and (ii) on the grain size d of the
material in question. In figure 25, the reductioW factor
is quantitatively presented.
As can be seen here, the non-linear flow effects are
negligible for materials with d in the range of fine and
medium sand, this even for hydr~ulic gradients up to I = 10.
or even more. But, if d is - say - 10. rom, the apparent
coefficient of permeabilIty under a gradient of I = 1 is
only about 1/10 of that according to Darcy. Fortunately, the
hydraulic gradients in drainage layers are normally in the
order of 0..1 or even less, but - as we see from the
diagram - a significant influence of non-linear flow effects
in coarse grained material still exists here.
This again should cause us to provide ample safety
reserves in the design of drainage layers in water storage
dams in order to provide "hydraulic safety" for them.
5. Concluding remarks

In the preceding pages an attempt has been made to present


various aspects of the filter problem in a clear form.
Besides pointing out the different filter rules and some
special aspects of their application, it seemed important to
draw our attention to the statistical nature of this problem
and to the fact that a thorough understanding and
analysation calls for a statistical approach.
As far as drains are concerned, we have shown the
limitations of Darcy's law and how the apparent permeability
decreases for coarse grained soils. The following brief
recommendations serve as concluding remarks:
266

Filters
Do not confuse filter and drainage functions in dams.
Keep the problem of autostability of base, filter and
drain materials in mind.
Apply filter rules to the safe border lines of gradation
bands.
Do not forget to provide enough filter thickness, this
with respect to the coarseness of the material in
question.
Collect data on the statistical scattering of gradations
of dam materials as a basis for further development of
statistical filter concepts and publish such data.
Try to apply statistical filter concepts and report on
experience.
Drains
Make a rational design of a drain with regard to its
capacity.
Do not forget restrictions regarding drain materials in
connection with their permeability.
References

Bertram, G.E. (1940) "An experimental investigation of


protective filters", Graduate School of Engineering,
Harvard University, Cambrigde/Mass. Pub. 267, Series
7.

Brauns, J. (1985) "Erosionsverhalten geschichteten


Bodens bei horizontaler Durchstromung", Wasserwirt-
schaft, 75, 10, 448-453.
Brauns, J. and witt, K.J. (1987) "Proposal for an
advanced concept of filter design", 9th European
Conference on Soil Mechanics and Foundation Engi-
neering, Dublin, Ireland, Discussion Paper 1988.

ICOLD (1990) "Conventional methods in dam construct-


ion", Bulletin 76, International Commission on Large
Dams, Paris.

ISSMFE (1987) "Report on Filters", (revised draft,


August 1987) Technical Committee on Filters of the
International Society of Soil Mechanics and Foundat-
ion Engineering.
Karpoff, K.P. (1955) "The use of laboratory tests to
develop design criteria for protective filters",
Proc. ASTM 55, 1183 - 1198.
267

Kenney, T.C., Chatral, R., Chin, E., Ofoegbn, G.I.,


Omanage, G.N. and Ume, A. (1985) "Controlling con-
striction size of granular filters", Canad. Geotechn.
Journal, 22, 32-43.
Leatherwood, F.N. and Peterson jr., D.F. (1954)
"Hydraulic head loss at the interface between uniform
sands of different sizes", Trans. American Geophys.
Union, Vol. 35, 4, 588-594.
Lubockov, E.A. (1969) "Berechnung der Suffosionseigen-
schaften nichtbindiger Baden auf der Grundlage des
Analogons ohne Suffosion", Kniznice odbornych a
videckych spisu Vysokelw nceni technickeno v Brne sv.
5, 135-146.
Prinz, E. (1923) "Handbuch der Hydrologie", 2. Aufl.
Julius Spinger, Berlin.
Sherard, J.L., Dunnigan, L.P. and Talbot, J.R. (1984)
"Basic properties of sand and gravel filters", ASCE
Journal of Geotechn. Eng., 110, 6, 684-700.

Sichardt, W. (1952) "Kies- und Sandfilter im Grund- und


Wasserbau", Die Bautechnik 29, 3, 72-76.
Terzaghi, K. and Peck, R.B. (1948) "Soil Mechanics in
engineering practice", John Wiley and Sons, New York.

Thanikachalam, V. and Sakthivadivel (1974) "Rational


design criteria for protective filters", Cando Geo-
techno Journal 11, 309-314.

US Corps of Engineers/WES (1953) "Filter experiments


and design criteria", Techn. Memo 3-360, Vicks-
burg/Miss.
Vaughan, P.R. and Soares H.F. (1982) "Design of filters
for clay cores of dams", ASCE Journal of Geotechn.
Eng. 108, 1, 17-31.
Witt, K.J. (1986) "Filtrationsverhalten und Bemessung
von Erdstoff-Filtern", Veraff. Institut fur Boden- u.
Felsmechanik, Universitat Karlsruhe, 104.

Wittmann, L. (1980) "Filtrations- und Transportphano-


rnene in porosen Medien", Veroff. Institut fur Boden-
u. Felsrnechanik, Universitat Karlsruhe, 86.
268

Wittmann, L. (1981) "Die analytische Ermittlung der


Durchlassigkeit rolliger Erdstoffe unter besonderer
Berucksichtigung des nichtlinearen Widerstandsgeset-
zes der Porenstromung ll , Veroff. Institut fur Boden-
u. Felsmechanik, Universitat Karlsruhe, 87, 115-142.

Ziems, J. (1968) "Beitrag zur Kontakterosion nichtbin-


diger Erdstoffe", Dissertation TU Dresden.

Zweck, H. (1959) "Versuchsergebnisse uber die Zusammen-


setzung von Filtern", Mitt. Bundesanstalt fur Wasser-
bau, Karlsruhe, 12, 15-28.

Zweck, H. und Davidenkoff, R. (1956) "tiber die Zusam-


mensetzung von Filtern", Mitt. Bundesanstalt fur Was-
serbau, Karlsruhe, 6, 24-33.
CHAPTER 11
STRESS - STRA IN LAWS AND PARAMETER VALUES
D. 1. NAYLOR

1. Introduction

Siress-sirai n laws for soil appear to have been developed with restrictions on their
applicability. if any, relating to whether the soil is cohesive or no n-cohesive rather
than whether it is insitu o r a fill . The same models have been used for both, only
the parameters are different. T his is not completely true. For example, a dedicated
model is used for insitu London clay, and late r in this chapler it will be noted that a
difficulty in app lying the critica l state model to fills is that it was o ri ginally developed
(or laboratory re-consliluled clays (Le . depositied in wate r) , Nonetheless, the
formulation of the basic models is not affected by this d istinction. Consequently much
o f the content of th is c hapter applies to insitu soils as well as to fills.
A stess-strain or consrirulille law can relate either the accumulated or the
inc remental stress to the corresponding strains. This can be ell pressed by:

(I )

Here d indicates increment. It can be interpreted as the increment from zero in


which case ~,d.f. become 2:,.f.. At the other elltreme the inc rement can be very
sma ll so that d approllimates the differential operator. 2:,.f. are vectors of the stress
and strain te nsor components. They can have length 6, 4 or 3 for respecti vely
3-dimensional , 3lIi-symme tric and plane strain formulations , e.g. for the last, 2: ::
(Ull.Uy,TlIyIT . Q is the corresponding mod ulus matrill, il is usua ll y symmetric.
Stress can be effective or total , and the Q matrill must correspond to th is. Where
it is necessary to distinguish between them a prime is added to idenlify effective, Le .
2:', Q.', Note Ihat for drain ed loading in which the pore pressure Cif any) remains
constant ~ :: ~' . The undrained case is considered in Chapler 12 , Section 8 whe re
a relationship between Q and .12' is de rived .
Equalio n (I) is central to the finite e lement method. This chapter is about its
form, i.e. how the components of the .12 matdll are defined . The Slress-S lrain la'Ml
described here for the most part re late e!!eclive stress 10 strain, and this may be
assumed for lac k o f a statement 10 the contrary. The prime will be omitted it being
understood thai slresses are generally effective. Jus t as c and 'fJ may define strength
in terms of effective or total stress so also it may be appropriate in some si tuations
to define the Q matrix components in te rms of total stress. This could arise with
partly saturated fil ls. It then becomes important to ensure thai loading rales and
drainage conditions in lab . tests correspond to the field conditions.

'69
E. ManltlM diJs NtYes (td.). AdwltlUJ il1 Roclcfil/ Slruc/lU"eJ. 269-290.
C 1991 KIIlWf'TAcadtmicPublisMr$.
270

The simplest form of the 12. matrix is linear elastic isotropic. Two elastic
parameters (e.g. Young's modulus, E, and Poisson's ratio, v, or bulk modulus, K, and
shear modulus, G) fully define the matrix. The 'transversely isotropic' linear elastic
assumption is the next stage up in sophistication. This features a vertical stiffness
different from the horizontal. Five elastic parameters are required. These elastic
models are well known and are not considered further here. (See, however, Section
8.1 of Chapter 12 for a convenient form of the isotropic elastic plane strain 12.
matrix.) Suffice it to note that it is good practice to precede a non-linear finite
element analysis by a linear one to get an idea of the results to be expected.
A stress-strain law should ideally incorporate the following characteristics of fill
materials:
(1) The increase in bulk stiffness which occurs with increasing stress level (i.e.
the concave form of an isotropic compression or oedometer stress-strain
curve(i.
(2) The reduction in shear stiffness which occurs with increasing deviator stress
(Le. the convex form of the triaxial test deviator stress-strain curve).
(3) A Mohr-Coulomb or similar type of failure criterion. This implies that an
incremental shear modulus tends to zero as the failure stress state is
approached.
(4) A higher stiffness on unloading.
(5) A higher stiffness at low stress levels (the threshold effect) and on
reloading following unloading.
(6) The effects of dilatancy, Le. the tendency of a stiff soil such as a well
compacted granular fill to increase its volume on shearing, or of a soft
clay fill to reduce its volume on shearing. Such a clay would be
negatively dilatant.
(7) Collapse settlement, i.e. the reduction in volume of initially unsaturated
material when saturated. This particular applies to rockfills.
The last of these should be considered separate from the rest. As is explained in
Chapter 12, Section 12 it can be catered for by providing a separate set of
parameters for whatever model is being used.
The models to be described here divide into variable elastic (Sections 2 and 3
below) and elasto-plastic models of which only the critical state model is considered
here (in Section 4). The former can incorporate characteristics (1), (2) and (3) and
to some extent (5) but not (6). (4) can be incorporated only by a 'user decision' on
what constitutes unloading and the alteration of the stiffness parameters for the region
in which this occurs. Elasto-plastic models of the critical state type can in principle
incorporate all of characteristics (1) to (6). How well they do it is, however, another
question!
All the models are in incremental form so that D in Equation (1) indicates a
relatively small increment. The 12. matrix is then 'incremental' or 'tangential'.
Following the formulation of each constitutive law an indication is given of how its
parameters can be determined. Also examples are included which give some idea of
their magnitude. Finally comments are made on the validity of the models.

(i) This is not always concave with rockfills. It may be convex due to a crushing
of particle contacts at higher stress levels.
271

2. Hyperbolic and Ec-KQ Models


2.1 BACKGROUND

These two models are appropriately described together as they have both been used in
a single analysis. This was done in the predictive analysts of the Beliche dam in
south-east Portugal by LNEC(i) (Naylor et ai, 1986). Each model represents different
parts of the dam.
The hyperbolic model is the best known of the early non-linear stress-strain laws
used in finite element analyses. It is attributed to Kondner (1963) who proposed that
the deviator stress-axial strain curve from triaxial tests should be approximated by the
hyperbola
f 1
(2)

in which a and b are constants for a particular test but in general will vary with the
cell pressure IJ 3. Duncan and Chang (1970) developed it into a full constitutive law
the formulation of which is given below. The generalised model involves at least
eight material parameters (nine if unloading is included).
The Ec-Ko model was developed at LNEC specifically for use in the central parts
of fills where conditions are approximately laterally constrained, i.e. they approximate
the 'Ko' condition. 'Ec' is the name of the constrained modulus. The model
involves a minimum of four parameters (seven if unloading is included).

2.2 HYPERBOLIC MODEL: FORMULATION

A tangential Young's modulus, E t , is defined by the slope of the (IJ 1 - IJ 3) E1 curve


constant). This is obtained by differentiating Equation (2) to give
(IJ 3

where
Rf (IJ 1 - IJ 3) (1 - sin 'P)
(3)
2IJ 3 sin 'P + 2c cos 'P

In this Pa is the atmospheric pressure, c and 'P are the Mohr-Coulomb strength
parameters, Rf is a reduction factor which reduces the f 1 = 00 asymptote of a- l-a- 3 to
the assumed deviatoric stress at failure, and K and n are constants to be determined
experimentally.
Poisson's ratio is defined by a separate hyperbolic relationship between the axial
and radial strain for the shear stage of the triaxial test of the form

E 3
E 1 (4)
F + DE3

(i) Laborat6rio Nacional de Engenharia Civil


272

Differentiation of this leads to the following expression for the tangential Poisson's
ratio

(5)

X is the same expression as in Equation (3), and G, F and D are further parameters
to be determined experimentally.
So far eight parameters have been defined to which values must be assigned: K, c,
'P, Rf, n, G, F and D. (This excludes Pa which will be assigned a nominal
atmospheric pressure in appropriate units, e.g. 100 kPa). If a stiffer response on
unloading is to be incorporated an increased value of K, namely KUf' must be
assigned.
A variation on the above formulation suited to fills with a curved Mohr-Coulomb
failure envelope is to set c=O and let 'P vary according to

0"
'P = 'Po - d'P Log ( ~ ) (6)

'Po and drp then replace c and 'P as parameters.

2.3 Ec-Ko MODEL: FORMULATION

Explicit expressions for the vertical constrained tangential modulus Ec and the
horizontal-to-vertical stress ratio Ko for conditions of no lateral strain are defined.
These constitute elastic tangential moduli and can be used to define E t and Pt.
(Equations (to) and (11) below.) The formulation is due to Veiga Pinto (1983).
The following relation between vertical stress and strain in the one-dimensional test
is assumed

(7)

in which Pa is the atmospheric pressure and Ac and Be are dimensionless constants to


be determined experimentally. Differentiation of equation (7), followed by re-use of
it to eliminate f" gives

dO",
(8)
df,

Ko(=O" /0",) is assumed to vary linearly with II, according to

Ko = Ao + B ~ (9)
o Pa

in which Ao and Bo are further dimensionless experimental constants.


Differentiation and rearrangement of Equation (9) gives a tangential Ko value

K t = ~ = Ao + 2Bo ~ (10)
o dO", Pa
273

The tangential Poisson's ratio and Young's modulus are obtained in terms of Ec
and Kot from elastic theory as

(11 )
1 + Ko t

(12)

The four parameters Ae, Be' Ao and Do cover loading. If there is unloading then
an additional three parameters are required. These comprise Co and Do to replace
Ao and Bo in (9) and (10) and thus define an unloading Ko ' and Ce to replace Ae
in 7 on the basis that Be = 1 for unloading. Thus Ec is replaced by the constant
CePa for unloading.

2.4 HYPERBOLIC AND Ec-Ko MODEL: PARAMETERS

The hyperbolic model parameters are normally obtained from consolidated drained
triaxial tests. A set of these are required at different cell pressures as would be done
to define a failure envelope. The Ec-Ko model parameters are obtained from
no-Iateral-strain compression tests in which the lateral stress (J 3 is measured.
The parameters are obtained by fitting the appropriate equations to the
experimental stress-strain curves. Details of this for the hyperbolic model are given
by Duncan and Chang (1970). For the Ec-Ko model a (J 1 : f 1 plot will yield the
parameters Ae,B e of Equation (7). A Ko = (J /(J 1:a- 1 plot is required to obtain Ao
and Bo '
As an example of the parameter values appropriate for an embankment dam the
core and inner shell values for the Beliche dam analyses described by Naylor et al
(1986) are reproduced in Table 1. In these analyses the Ec-Ko model was used for
the core and foundation and the hyperbolic model for the rest of the dam.

TABLE 1. Beliche dam predictive analyses, LNEC parameters

Hyperbolic model material parameters

Zone K n 'Po:deg 4p:deg Rf G F D

Inner shell 900 0.25 49.9 13 .4 0.89 0.28 0.18 5.5

Ec -Ko material parameters

Zone Ae Be Ao Bo

Core 6805 2.17 0.62 -0.0092


274

3. K -G Model

3.1 BACKGROUND

The idea behind this model is that it should incorporate as a first priority
characteristics (1) and (2) above. With an isotropic elastic formulation this implies
that the incremental, or tangential, bulk modulus, K, should increase with increasing
mean stress, and the shear modulus decrease with increasing deviator stress. It also
incorporates a yield criterion (characteristic (3)).
The model was conceived as an alternative to the hyperbolic model as it was felt
that in view of the limitations of variable elastic models it was inappropriate that as
many as eight or nine parameters should be required. Could not nearly as good
results be obtained with five parameters?

3.2 FORMULATION

The basic model may be defined in either of two ways. The first, referred to as the
(Tm,(Tq K-G model defines K and G in terms of the stress invariants (Tm and (Tq
where

(Tm (13 )

(14)
and the second, the (Ts,(Td K-G model, defines K and G in terms of the stress
invariants (Ts and (Td where

In these z is the out-of-plane direction. For plane strain applications (for which the
(Ts,(Td model is particularly appropriate) it is convenient to use K = K + 1/3G since
this defines the slope of the (Ts:volumetric strain curve for plane strain (as does K the
(Tm:volumetric strain in general). The definitions are as follows:

(Tm.lTq K-G model

K Klm + aKffia- m (17)

e elm + aeffia- m + {leffia- q (18)

ITslTd K-G model

K Kl + O'KlTs (19)

e e l + ae(Ts + {le(T d (20)


275

In these the five parameters (K,m, ... ,I3G m or K" ... ,I3G) are material constants. I3G
is negative. The superposed m is used to distinguish the (f m.(f model from the
other. No tag is put on the (fs,(fd parameters as these are most ~requently referred
to in what follows.
For unloading higher values of these parameters are required. A simple assumption
is to increase all five parameters by a single factor (typically in the region of four
times). This implies that Poisson's ratio is the same in unloading as in loading.
The (fs,O"d version is particularly appropriate if it is assumed that the soil obeys a
Mohr--coulomb criterion. This criterion is incorporated in this version of the model
as will now be demonstrated. At yield the tangential shear modulus G becomes zero.
Setting G=O in Equation (20) and rearranging gives

(21)

The Mohr--coulomb yield criterion can be written in the form

(22)

Comparing (21) and (22) gives

2 sin 'P (23)

~ ~ 2c cos 'P (24)


-l3c
These relations help in assigning values to O'(}.I3G and G,.
By a similar argument it can be shown that the O"m'O"q version incorporates a
Drucker-Prager or extended Von Mises type of yield criterion (i.e. a cone in principal
stress space).

3.3 K-G PARAMETERS

K, m and Cl'J(m may be determined directly from an isotropic consolidation test. The
slope of the O"m: (v (v = volumetric strain) curve is K. and this may be plotted
against O"m. The best straight line fit will define K,m and Cl'J(m.
The shear modulus parameters may be found by noting that Equations (18) and
(20) define a plane in respectively a G,O"m'O"q space and a G.O"s.O"d space. The line
formed by the intersection of this plane and the G=O plane identifies a failure
condition, i.e. Mohr--coulomb for O"s,O"d version and a Drucker-Prager type for the
O"m,a q version. Parameters G,m, O'(}m, I3G m or G" aG, I3G may therefore be
determined by plotting the results of conventional triaxial tests as points in a G ,0" m,a q
or a G,as.ad space respectively and finding the plane which gives a best fit to these
points.
An example of this plane fitting procedure is illustrated in Figure 1. This is based
on three consolidated drained triaxial tests on a boulder clay which was used for the
core in an embankment dam. The spot values on the inclined lines are the G values
obtained by careful measurement of the slope of the stress strain curves (i). The best
fit surface is indicated by the G-O, 10 and 20 MPa contours. This surface has the

(i) These relate O"d to (d where (d = (,-1/3(v. The slope of this curve is 3G.
276

equation

G 500 + 27(fs - 27(fd kPa (25)

whence G 1 = 500 kPa, 0'(} = 27, {3G = -27. Equations (23) and (24) show that
these values imply that <p = 30' and c = 21 kPa.
The K-G model stress strain curves corresponding to the three tests can then be
calculated using Equation (20). They are compared with the laboratory curves in
Figure 2. The (fd:fd curves agree remarkably well. A weakness of the model,
however, is revealed in the fV: Ed curves. It underestimates the compressive
volumetric strain. It cannot incorporate the negative dilatancy of this soil (i.e.
characteristic (6) above).
It is somewhat time consuming to carry out the plane fitting procedure. An
alternative is to decide on c and <p and then select the three parameters so that
Equations (23) and (24) are satisfied. This leaves one independent parameter which
can be selected to give the right shear stiffness, the strength requirements having first
been satisfied.
Oedometer tests may also be used to assist in determining the five K-G
parameters. This, however, is more difficult as it is not possible to write an explicit
equation relating (f 1 to E 1 for the no-lateral-strain case. This can only be obtained
by an incremental analysis. Also it involves all five parameters. However, if the
three shear modulus parameters have already been determined then by a trial and
error process oedometer test results can be used to select values of the two bulk
modulus parameters (K 1m, CXJ(m or 1<1' a).
The K-G parameters used for the core and inner shell in the Beliche dam
prediction analyses are given in Table 2. The 'prediction' (P) values correspond to
the hyperbolic-Ec-Ka model parameters of Table 1. The resulting settlement profiles
are compared in Figure 23 of Chapter 12. Two other sets of K-G model parameters
are included in Table 2. These were used in a subsequent back analysis of the
Beliche dam which attempted to reproduce the measured settlements of the dam.
Settlement profiles from these are reproduced in Figures 24 and 25 of Chapter 12.
The 'dry' parameters relate to the fill as constructed and the 'wet' to it after it had
become saturated during impounding. (This applied to the upstream shell only.)
These two sets were used to model collapse settlement. Triaxial and one-dimensional
stress-strain curves obtained from these parameters for the inner shell are compared
both with critical state model curves and experimental in Figure 7 below.
The large differences between the parameters of Table 2 for the inner shell
material are noted. This is due partly to the fact that the material did not fit the
model very well. A good fit to both triaxial data and one-dimensional tests could
not be obtained. There was, therefore, an element of subjectivity in selecting the
parameters. Another factor was the variability of the material. It was difficult to
know how representative were the tested samples. The prediction parameters
overestimated the stiffness, thus all the parameters (except K 1) have been considerably
reduced for the later analyses. Note, however, that the strength is little altered.
The changes to the core parameters (and also those for the outer shell which are not
presented here) were much less drastic.
277

...
E
z 500
.:x:

o L---~-----L~----~------~---------L--__ aS
o 500 1000
kN/m2
Contours are of best fit surface
Spot values are G in MN 1m2

Figure 1. K-G model fit to triaxial test shear modulus values (after Naylor et ai,
1983)

TABLE 2. Beliche dam K-G parameters


Key: P = Prediction analysis (dry) (see Naylor et ai, 1986)
D = Dry back analyses
W = Wet, i.e. post-saturation, back analyses

K, MPa C, MPa QK ClC {3c c*MPa <p*

Clay core (P) 5.7 5.0 68 22 -34 78 19


Clay core (W)t 2.5 4.5 56 19 -30 79 19

Inner she II (P) 15.1 13.3 210 176 -138 63 40


Inner she II (0) 16.0 2.2 56 41 -30 50 43
Inner she II (W) 3.3 1.0 56 17 -15 40 35

* Implied by G" ClG' {3G


t The core was assumed to be placed wet.
278

o
~
........::

0'] =
----
520 kPa
O'd #
kPa
Legend
500 500
Experimental----
K-G-model - - -

10
o O~~~L-L-L---_-----~I--~~

Figure 2. K-G model fit to triaxial test stress-strain curves

3.4 K-G MODEL - AN ALTERNATIVE

Alteration of Equation (20) to

G (26)

allows the De Mello form of yield surface described in Chapter 5 to be incorporated


in the K-G model. K (or K) is defined as previously. In Equation (26) b is an
additional parameter between 0 and 1, and Pa is the atmospheric pressure (introduced
so that D'G continues to be dimensionless). Substitution of b=l causes (26) to revert
to (20). For b less than one G 1 would normally be set to zero.
The G contours of Figure 1 become curved if b is less than one. In particular
the G=O contour, in common with the De Mello yield line, has a vertical tangent at
O"s=O
279

4. Critical State Model

4.1 BACKGROUND

This model originated from work done at Cambridge University, England, in the 1950s
and 60s under the leadership of Professor Roscoe. His objective was to develop a
theoretical framework which incorporated the essential characteristics of, at least, clay
soils. The model is more than just a stress:strain law. It has become a philosophy
providing a means of communicating ideas and of teaching soil mechanics. Atkinson
(1983) elevates it to the status of a paradigm(i).
The model unifies the previously unconnected concepts of shear strength and soil
deformation properties. It brings together the accepted concepts of critical state (or
critical voids ratio), voids ratio :Log. of stress relations for clay soils, irreversibility of
displacement on unloading, and the Mohr-Coulomb and Hvorslev strength criteria.
There is not space here to enlarge on this and the interested reader is referred to
the original text by Schofield and Wroth (1968) (unfortunately now out of print), the
texts by Atkinson and Bransby (1978) or Britto and Gunn (1987), or Chapter 7 of
Naylor et at (1983).
The model can be viewed as a form of strain hardening plasticity. In contrast to
metal plasticity where strain hardening (if any) is related to deviatoric plastic strains
(Le. irrecoverable distortion) the strain hardening in the critical state model is related
to volumetric plastic strain. The formulation given here is on the same lines as for
conventional elasto-plasticity with this principal difference.
The suitability of the critical state model for fills must be questioned. It
incorporates the properties of clay soils, in particular the one to one relationship
between voids ratio and effective stress of a clay soil as it is consolidated from a
slurry. A relationship of this sort will have some validity for a soft clay fill but little
for rock or sand-gravel fills. Nonetheless other characteristics incorporated in this
model are present in fills, and it does seem that it can provide a better
representation of at least some fills than the variable elastic models. This particularly
applies when effective stress unloading applies, as in the impounding behind both
central core and upstream membrane dams.
There are many possible variations on matters of relative detail in the formulation.
Some of these may be important in particular applications. Certain of these
variations will be considered after the basic model has been described.
The formulation given here features the following
(1) The yield function is described in terms of the 'plane strain' stress invariants ITs
and lTd. This means that the Mohr-Coulomb yield criterion is a natural part of
the formulation. (The yield surface takes the angular form in principal stress
space shown in Figure 3.)
(2) Linear elasticity is assumed within the yield surface.
The original formulation differed on both these points. The yield function was
prescribed in terms of 'p and q' (ITm and IT , respectively in the present notation).
Also variable elasticity was prescribed. A furt%er, more detailed difference is the use
here of a parameter X to describe the plastic bulk compressibility. This represents a
simplification of the conventional formulation.

(i) This word is used in a different sense to that given in some dictionaries. It
implies something bigger than a theory - a framework of ideas within which
theories can be developed. It is essentially open ended.
280

Figure 3. Critical state model in principal stress space

4.2 BASIC FORMULATION

The first part of this is quite general and applies to any form of strain hardening
plasticity. Let

F(![,h) =0 (27)

define the yield function. h is a hardening parameter to be defined later. Equation


(27) represents a set of closed surfaces in stress space (one for each value of h) on
or within which the point representing the current stress must lie. Also let

Q(Q:,h) =0 (28)

define a plastic potential. The gradient of Q defines the flow rule vector !!q, i.e.

aQ (29)
~q
Oil
Also define

=
aF (30)
~f
Oil
For associative plasticity Q=F (and !!f=~=!!). For generality it is assumed that Q
and F differ and then later for certain conditions they can be equated.
The flow rule is used to inter-relate the components of plastic strain increments(i)
according to

(31 )

(i) An alternative is to use plastic strain rates ~ P = d f Pldt. Through multiplication


by dt changes this back to the incremental formulation.
281

d}" is sometimes called the 'plastic multiplier'.


Equation (31) only operates when plastic yielding occurs. The stress is then on the
yield surface and Equation (27) applies. Figure 4 illustrates the situation. In this
representation the plastic strain components of Equation (31) are the components of
the vector normal to the plastic potential if the axes represent both stress and plastic
strain increments. Thus a plane strain situation could be represented by three
orthogonal axes representing both O"X 'O"y , T and d EXP ,d fyP ,d)'P.
A parameter H is now defined by

Hd>- = !!(T d![ (32)

H has the dimensions of stress. In simple plasticity formulations it is a material


property which defines the plastic stiffness. It is a variable for the critical state
model and its form will be derived later on.
A basic assumption of elasto-plastic models is that strain is the sum of elastic and
plastic components. Therefore

(33)

Noting that d.fe ~ -ld![ where ~ is the elastic modulus matrix, and introducing
(31), gives

(34)

Pre-multiplying by ~fT~ and noting that ~~ -1 is the identity matrix gives

!!(T~d.f = !!(Td![ + ~fT~d>-!!q

Eliminating !!(T d![ by (32) and re-arranging gives

1
d>- = -
(3 -afTD dE
-e- (35)

in which (3 =H + !!(T~!!q

Multiplying (34) by ~, eliminating d>- by (35) and re-arranging gives

1
d![ = ~ed.f - ~ (~e~q)(~fT~e)d.f
Noting that ~fT~ = (~~f)T one obtains

d![ = ~p d.f (36)

with
1
~ep = ~e - .Qq.QfT (37)
~
in which !?q = ~!!q
and !?t = ~~f
282

J~q referred to f! axes


/l~~d~P referred to d~P axes

Origin of
stress--~
axes

Figure 4. Yield surface and plastic potential with flow rule

< Flow rule ':x cs l


~1 ,// '\
State boundary surface

o
o "'-~:::-_-+--_ _ _---L_"O" s
I. 0"( .1
(b)
h

D~---~-~

o~--~--~------~~
LnO" s

(a)

Figure 5. State boundary surface showing typical yield surface


283

Equations (36) and (37) define the constitutive law for elasto-plastic models in
general. There is one further step of general validity to take before it is necessary
to consider the special conditions of the critical state model. This is the formulation
of the consistency rule.
The consistency rule is obtained by differentiating Equation (27) and setting the
differential to zero, i.e.

dF(~,h) = ~.~ +
aF aF
Oil.ah = 0

Introducing (30) and eliminating !\.fT d~ by (32) gives

aF
Hd)" + Oil dh = 0 (38)

Equation (38) is a mathematical statement that the stress point sticks to the current
yield surface during yielding noting that this yield surface is itself changing in size on
account of the variation of h.
It is appropriate now to introduce some of the jargon of critical state soil
mechanics. The terms 'state boundary surface' (s.b.s) is used to define the set of
yield surfaces of Equation (27). This is illustrated in Figure 5 as a surface in a
i'Ts,i'Td,h space. Its intersection by a h = constant plane defines the yield surface (a
line in this representation) for that particular h. This is referred to as an 'elastic
wall'. Conventionally the s. b.s. is represented in a space with the axis h replaced by
the voids ratio, e. (The relation between hand e need not concern us here. It is
not one to one since e contains elastic as well as plastic components.)
h is defined as the volumetric plastic strain, i.e. h = fxP + fyP + f zP. Using
Equation (31)

(39a)

or for plane strain situations in which EZP o(i).

dh = d>- aQ (39b)
00s
Substituting (39) into (38) and dividing through by d).. gives, generally

H
aF (40a)
-Oil

or for plane strain

(40b)

The shape of the yield surface, and more generally the shape of the state boundary
surface as this includes the specification of how F varies with h, must now be
prescribed. So also must a plastic potential Q. This will first be formulated for the

(i) This is for the Mohr Coulomb formulation, but will not generally be the case.
284

plane strain case so that use will be made of Equation (40b) rather than (40a).
Figure 5 shows the general shape of the s.b.s. Its shape is based on experimental
findings(i). All the yield surfaces are similar in shape and their size can be
prescribed by the position of C, i.e. by CJc in Figure 5. The soil is at the critical
state when its state is represented by point C. Its characteristic then is that there
must be no volumetric strain change. The soil at this point will deform at constant
stress. This has implications for the flow rule which must prescribe zero plastic
volumetric strain at C. This means that for plane strain the flow rule arrow if
represented in a CJd,CJs space must be parallel to the CJdaxis.
The change of CJ c with h is assumed to be exponential, i.e. the 'critical state line'
which is the locus of the point C is a straight line in the semi-log plot of Figure
5(c). The slope of this line is the material parameter X. It is defined as follows (as
can be seen from Figure 5b).

dh = L
CJ dCJ ( 41)
e C

X is related to the loading and unloading compression index parameters l\ and K


which are used in the references given above by

A-K (42)
X l+e

Equation (41) is the hardening law. As F will be expressed in terms of CJc rather
than h it will be convenient to use (41) to make the substitution

(43)

The conventional formulation assumes a slope discontinuity in the s.b.s. at the


critical state line. A different formulation for F and 0 is therefore required either
side of C. The region to the left of C will be called the over-consolidated (o.c.)
region and that to the right the normally-consolidated (n.c.) region. In a CJd:CJs
stress space a straight line yield surface is assumed for the o.c. region and an ellipse
for the n.c. The flow rule will be associative for the latter, i.e. Onc :: F nc' but
non-associative for the o.c. region. The shape of the n.c. yield surface and o.c.
plastic potential are such that the flow rule will give no plastic volume change (and
as a consequence of this zero strain hardening) at the critical state point, C. The
formula for F and 0 are as follows

o.c. region
(44)

in which S :: S/Scs and Sand Scs are respectively the slope of the yield line and
critical state line, as shown in Figure 5(b).

(i) In the original Cam clay model developed at Cambridge the shape was derived
from theoretical consideration rather than as a direct fit to test data.
285

QOC ud + Ro (U C S )2 = 0 (45)
= 2U-U
c

in which Ro = RalScs and Ro = 2 Sin "'0 '"


is the dilatancy angle and "'0
its
assumed value at Us = O. The form of Equation (45) causes '" to vary smoothly
from zero at C (when Us = u c ) to "'0
at the origin thus satisfying the critical state
requirements. Generally "'0
will be less than <p.
Equation (44) is a Mohr-Coulomb yield criterion in disguise. It can be arranged
into conventional form by making the substitution S = 2 sin <p and (Scs - S)u c = 2c
cos <p. This gives

(46)

which is the same as the form used with the K-G model. i.e. equation (22) above.

n.c. region

(47)

It is easily verified that at the critical state (when Us = U c) F nc = F oc and also that
Onc prescribes zero plastic volume change. i.e. '" = O.
Determination of the components of !2ep. i.e. (:1. ~. l!.f involves first differentiating
F and 0 with respect to Uc and the stress invariants us.ud. Then the partial
derivatives of Us and ud with respect to the plane strain components of stress are
required. Details are given in Chapter 7 of Naylor et al (1983).
A general form of the model suitable for three dimensional applications can most
easily be obtained by proceeding exactly as above but using Equation (40a) instead of
(40b) and um.uq in lieu of us.ud. This formulation has the disadvantage that the
o.c. part of the model does not incorporate a Mohr-Coulomb criterion. Just as for
the um.uq K-G model it incorporates a Drucker Prager type of yield surface. If
depicted on Figure 3 it would be circular in the 'Pi' plane and would enclose the
Mohr-Coulomb surface coinciding with it along the lines A.
A formulation suitable for 3D use incorporating the Mohr-Coulomb, or indeed any
other preferred yield criterion, may be obtained by defining F and 0 in terms of urn'
U and e. where e is the Lode angle which measures the orientation in the Pi plane.
T'l.is involves some algebraic complexity but does not introduce any new concepts.

4.3 VARIATIONS ON THE THEME

The writer has made a number of modifications to the basic model described above to
make it better for practical applications using the finite element method. There are
two main changes. as follows.
Firstly. the o.c. yield surface has been changed from the Mohr-Coulomb straight
line to an ellipse. Initially this was a continuation of the n.c. ellipse but this has
been changed to a 'flattened' ellipse. A 'flattening parameter'. p. can be assigned so
that if p.=O the M.C. line is retained. or if p.=1 the n.c. ellipse continues into the
o.c. region (Figure 6). Associative plasticity is assumed throughout. This has certain
computational advantages. p.=~ is probably suitable for most situations. In addition.
provision has been made to shift the yield surface and the critical state line to the
286

left by a constant crt, as shown in Figure 6. crt would normally be small.


The second change is more fundamental. It is a modification to the model to
allow some plastic yielding to occur while the stress point is still inside the yield
surface. This is believed to be more realistic as it avoids the sudden change in flow
conditions which occurs when the stress point first reaches the yield surface. The
generally smoother behaviour lends itself well to incremental analysis. It is referred
to as the continuous plasticity critical state model (c.p.c.s.m.) and is described by
Naylor (1985). Results from c.p.c.s.m. analyses are generally not very different from
analyses using the conventional formulation.

0' d oc n.c region


~.---------~--------~~~
/'

~ 1.4.
1- - - - ' - - - 1
. .

Figure 6. Modified form of c.s.m. yield surface

4.4 C.S.PARAMETERS

The following 7 parameters are required for the basic model described in Section 4.2:
Elastic moduli - E and v (or K and G)
Plastic compressibility - X
Slope of critical state line - Scs (or 'Pcs since Scs = 2 sin 'Pcs)
Slope of critical O.c. yield line - S (or 'P since S = 2 sin 'P
Dilatancy at zero stress - Ra (or fo since Ra = 2 sin fo)
Initial size of yield surface - (Jco
The model modified as described in 4.3 above dispenses with the parameters Sand
Ra (or 'P and ifo) but adds (Jt (or ccs = crt Tan 'Pcs) , I'- and a power index n needed
for the c.p.c.s.m. formulation. I'- and n, however, can be c1asssified as parameters
which determine the version of the model rather than material properties. They may
be fixed without reference to the particular situation under study. (Jt provides a
means of refining the model for low stress conditions and will often be zero.
There are therefore five essential parameters: E, J', X, Scs (or 'PcS> and (Jco' The
last three of these are normally the most critical. X may be found from oedometer
or other consolidation tests which measure the loading and unloading Compression
287

indeces(i). These, plus the voids ratio, e, can be substituted into Equation (42) to
give X although this tends to overestimate it. <Pcs may be found as the slope of the
post-peak strength envelope from drained triaxial or shear box tests. (Tco is not so
easily evaluated for fills. For insitu clays it is related to the pre-consolidation
pressure, and this can be estimated from oedometer tests. For fills one would expect
it to be related to the compactive effort, but it is not yet established how this should
be done.
Table 3 presents preliminary values for the Beliche dam core and inner shell.
Figure 7 shows stress strain curves for a triaxial and one-dimensional test on the
inner shell material obtained from these parameters. These curves are compared with
the experimental results and the curves obtained from the K -G model using the
parameters identified by (D) and (W) in the last two rows in Table 2.

TABLE 3. Beliche dam inner shell c.p.c.s.m. parameters (preliminary)


Model parameters n=2, p,=O.5. D,W as in Table 2

E MPa v ccs kPa <Pes X (Tco* kPa

Core (W) 25 0.3 25 25" 0.006 25


Inner she II (D 45 0.3 0 40 0.006 50
Inner shell (W 20 0.3 0 35 0.012 25

* This is the minimum value of (Tco. In the modelling of triaxial tests


(Tco is increased if otherwise the initial stress would lie outside the
initial yield surface.

(i) i.e. C c or Cs in conventional soil mechanics notation. These are related to ),.
and K by C c = 2.3),. and Cs = 2.3K.
288

1500,.-----r---,-----,----r-----,

Iii 1000
a..
..lr::
,.,
---
b
I
b
500

- - Laboratory
- -[,5, model
- - - K-G model
2 4 6 8 10
e: (%)

Triaxial test 0'3= 300kPa

1000

800

a.. 600
IV /' 1 Dry
::: //
b
400 / / End of .;}

200
... :::::.- ,..,
~ experiment ,.;.;
..::::-
Wet

0
0 2 3 4 5
e:'(%)
One dimensional test (e: 3=0)

Figure 7. Comparison of laboratory test, K-G and C.s.m. stress-strain curves for
Beliche dam soft rockfill
289

5. Conclusions

Variable elastic stress-strain laws are attractive in that they retain the simplicity of
linear elasticity and yet offer significant improvements on it by being able to
incorporate reasonably well the first three requirements listed in the Introduction.
Against this they cannot incorporate dilatancy (requirement 6), and they can only
incorporate unloading stiffness in a crude way.
A more subtle disadvantage of variable elastic laws compared with plasticity based
laws concerns the pattern of deformation during yielding. Measurement of strains in
soils (principally sandy soils) suggests that they do deform in the manner predicted by
elasto-plastic models. Basically this means that the deformation pattern (i.e. the
direction of the strain increment vector in strain space) depends on the accumulated
stress (Le. the point in stress space) and not the stress increment as assumed in
variable elastic models. This difference is not revealed when there is no rotation of
pincipal stresses as in a triaxial test. It becomes significant, however, when there is
significant rotation as in the development of a shear failure surface in an embankment
dam.
The results of variable elastic analyses should be interpreted in the light of the
above. The advantage of the hyperbolic model is that the procedures for determining
its 8 or 9 parameters are well established. But against this there are too many of
them. The K-G model is relatively simple with only 5 parameters but fixing their
values is somewhat 'messy'.
A difficulty encountered in the selection of parameters for the Beliche dam rockfill
is worth mentioning here as it may be representative of at least low quality rockfills.
It was the impossibility of finding parameters which matched both isotropic
compression tests (on triaxial samples) and one dimensional compression tests. The
former tended to be concave, i.e. of low stiffness at low stress, and the latter convex
exhibiting a relatively high stiffness at low stress. Parameters matching the isotropic
test would grossly under-estimate the one-dimensional and vice versa. The problem
applied to all the models described here. A possible explanation is that the dilatancy
featured in these soils at low stress levels gives them a high stiffness when they are
constrained laterally. This needs to be investigated further.
A considerable amount of experience has now accumulated of the use of the three
variable elastic models described here in the back analysis of the construction of
embankment dams, plus a little experience of the post construction stages. For this
reason alone they have some future in predictive analyses, at least for the construction
stage.
Elasto-plastic models of the critical state type have the greatest potential. In
addition to their theoretical advantages their parameters are for the most part readily
obtained from standard strength and compression tests. Little experience has as yet
been accumulated in applying them to fills. A particular matter which requires
investigation here is how to select the stress history parameter (Tco'
290

References

Atkinson, J.H. (1983)


'Is soil mechanics in a critical state?', Ground Engineering, 16, No. 1 (Jan.
1983), 2.
Atkinson, J. H. and Brandsby, P. L. (1978)
The Mechanics of Soils, McGraw-Hill, London.
Britto, A.M. and Gunn, M.J. (1987)
'Critical state soil mechanics via finite elements', Ellis Horwood, Chichester,
England.
Duncan, J.M. and Chang, C-Y. (1970)
'Non-linear analysis of stress and strain in soils', Proc. Am. Soc. Civil Engrs.,
96, No. SM5, 1629-1653.
Konder, R. L. (1963)
'Hyperbolic stress-strain response: Cohesive soils', Proc. Am. Soc. Civil Engrs.,
89, No. SMl, 115-143.
Naylor, D.J. (1985)
'A continuous plasticity version of the critical state model', Int. J. for Num.
Meth. in Engng., Vol. 21, 1187-1204.
Naylor, D.J., Maranha das Neves, E., Mattar, Jnr., D. and Veiga Pinto, A.A.
(1986)
'Prediction of construction performance of Beliche dam', Geotechnique 36, No.
3, 359-376.
Naylor, D.J., Pande, G.N., Simpson, B. and Tabb, R. (1983)
'Finite Elements in Geotechnical Engineering', Pineridge Press Ltd., Swansea.
Schofield, A.N. and Wroth, C.P. (1968)
'Critical State Soil Mechanics', McGraw-Hili.
Veiga Pinto, A.A. (1983)
'Structural Behaviour Predictions of Rockfill Dams', Thesis, L.N.E.C., Lisbon.
CHAPTER 12
FINITE EL EMENT METHODS FOR FILLS AND EMBANKMENT DAMS
0 . 1. NAYWR

1. Introduction

This chapter covers a ll types of fills which are built up in layers. This includes road
embankments as well as e mba nkment dams. Concern will be principally with the
latter as it is for the more complex fill s such as these that finite element analysis
becomes necessary.
There are two ma in types of embankment dam: that containing a relatively
im pervious internal core supported by rockfill shoulders, and the upstream membrane
dam. Usually rockfill and clayey fill are involved in the former and just rockfill in
the latter. The upslream membrane is usually e ither concrete or ashphalt.
An important purpose of finite element analysis of e mbankment dams is to
dete rmi ne internal st ress distributions . This is necessary to find out if there is a
danger of tensile effecti ve stresses (i.e. total stresses less than the pore pressure)
developing which could cause cracks. The analyses are also used to ca lculate
displacements. Although disp lace me nts are , at least in the case of clay core dams,
ultimately of less interest than the stresses they a re important because they ca n
relatively easily be measured . This makes it possible to calibrate a n analysis against
measurements laken al an early stage, or on a tr ia l fill, so that the accuracy of the
assumed soil stiffness can be improved . This in turn will lead to a more accurate
stress calculation.
The construction stage, covered in Sections 2 10 7, is the first and most
straightforward part of the overall process, and there is considerable experience of the
fin ite eleme nt mode lling of it . This started with work done principally in the U.S. A.
in the 196Os. Some of the ke y papers covering this and later work are given in the
Bibliography. The subsequent impound ing and operation stages, dea lt wi th in Section
9 el seq., are much more in the developme nt stage . A correctl y pe rformcd
construction analysis is a necessary prelude to the later analyses as they will carryon
from it.
Where pore pressures a rc involved the re is usually a choice between carrying out
analyses in terms of tota l or effective stress. In the case of dry rockfill the two
methods become the same. For clay materials, however , there is a distinctio n. This
has little bearing on the construction techniques described in Sections 2 to 7 but has
on the und rained construction of clay fl11s and on the post-constr uction ana lyses.
Effective stress techniques are t herefore described before the latter, i.e . in Section 8.
291
E. Marallha das Nel't's led. }. Alimllces ill k ()c~fi/l SI",,.mres. 291-340.
1991 Klmwr Acudrmic P,,/'/ishus.
292

It is shown that they have certain advantages due to their flexibility and the ease
whereby measured pore pressures can be incorporated in the analysis.
The techniques for modelling construction are only to a small extent affected by
the material law used to describe the soil. Material laws are covered in Chapter 11.
The Chapter ends with case studies which illustrate the modelling of construction
and the subsequent operation of three embankment dams.

2. Number of Layers - Actual and Analytical

Fills for road or rail embankments or for embankment dams are invariably built up in
compacted layers. These range in thickness from about 150mm for clay fills to about
1m for rock fills.
The thickness of the physical layer will, therefore, generally be small compared
with the height of the completed fill. In a typical embankment dam, say 50m high,
there will be of the order of 100 layers. It will not be practical to make the finite
element layers of similar thickness to the actual ones. There would be far too many
finite elements. Instead thick finite element layers (perhaps 10m thick) will be used
so that there may be as few as five.
The following four sections show how the finite element analyses can be carried out
and the results interpreted so that these thick analytical layers can correctly represent
the real situation. There are two issues here:
(1) the correct interpretation of computed displacements,
(2) the appropriate stiffness of a finite element layer as it is placed.
Before proceeding further deformation in a rising fill must be defined.

3. Deformation in a Rising Fill

When dams are built markers to measure settlement, lateral movement, or both are
installed within the dam. These are installed as the fill is being raised so that they
are embedded just below the current surface. The datum for movement is then
established, i.e. the deformation is zero when the instrument is installed. (Ideally it
should be on the fill surface but in order to prevent damage to the marker from the
compaction equipment it will usually be 1 to 1m below the surface.)
Deformation is therefore defined as the movement of a marker installed on the
surface of the rising fill. With this definition computer deformations can be directly
compared with the measurements.

4. Basic Finite Element Procedure

The basic procedure is quite straightforward and is illustrated in Figure 1.


A succession of finite element analyses are carried out with the loading for each
one comprising the application of gravity to the new layer. Each successive analysis
inherits stresses and displacements from its predecessor and these are accumulated.
The displacements, however, are accumulated in a special way as will be explained.
The finite elements in the layers not yet placed can either be omitted and added to
the mesh as they come into existance, or they may all be included from the start and
simply given a very low stiffness while they are in the 'ghost' category. This latter
procedure has the advantage of simplicity, but the alternative of adding the elements
293

as they are needed involves less computing at the expense of added complication.
The first stage in the analysis is to make the necessary geometric idealisation, Le.
whether the analysis is to be two- or three-dimensional, whether or not the
foundation is to be included in the mesh, and what type of elements should be used.
A factor to be considered here is the construction sequence. Is the fill to be raised
evenly so that layer boundaries can be assumed horizontal? Or is part of the fill
(such as an upstream cofferdam) to be constructed ahead of the rest? The layer
boundaries should be chosen to accommodate this, i.e. they should roughly correspond
to the actual fill surface as the fill is raised. A finite element layer, although it
cannot be less than one element thick, can (and often does) contain more than one
layer of finite elements.
Then the material properties must be chosen. This involves the selection of a
material model and the major decision of selecting parameter values. Provision must
also be made for the stiffness adjustment of the new layers. See Section 6 below.
In general initial stresses must be prescribed. These must be self equilibrating, and
if the foundation is included in the finite element model insitu stresses should be
assigned to it(i). If the ground surface is level these can be prescribed directly as )'h
vertical and Ko)'h horizontal (Ko generally having to be assumed). Otherwise a
preliminary excavation analysis from a level ground surface may be required to
calculate a self equilibrating stress field.
The initial stresses in the embankment itself will usually be set to zero - at least
in terms of total stress. This is consistent with the concept of a new layer being
initially weightless. The applied gravity imposes the stresses. This does not, however

11} 1
'Ghost' elements New layer at least
//~-,-", one element thick

L ~
Old I,ye"
I'~-I~
~=s..:ct",1 I,ym
{"'-Elements in foundation If compressible-------./'

Figure 1. Finite element modelling of fill construction

(i) These may be omitted if the ground is modelled as linear elastic as its stiffness
is not affected by the stress level. The stresses, if needed, can then be
determined later by adding an initial stress field.
294

take into account any stresses locked into the material by the compaction plant. This
will be considered in Section 7 where it is shown that for clay soils analysed in terms
of effective stress an initial negative pore pressure (suction) balanced by an equal
compressive effective stress (thus making the initial total stress zero) is required.

5. Interpretation of Finite Element Displacements - 10 Case

The one-dimensional case in which there is no lateral strain is a good approximation


of the physical situation in the early stages of construction of a large fill. It is used
here to develop a criterion for the interpretation of deformations computed in a finite
element analysis (with its inevitably unrealistically thick layers). It is also shown that
the thickness of the finite element layer is irrelevant in the 10 case in that the
stiffness assigned to the new layer has no effect. (This is not the case in 2 and 3D
as is explained in Section 6.)
Consider a fill made of a linear elastic material. It is supposed that in the real
situation layers are very thin so that the fill raising is effectively continuous. Figure
2 shows an example of a 6m high fill modelled by 3 layers of quadratic finite
elements. (As this is aID problem these elements could be simply 3-noded vertical
'string' elements.) The final stiffness, 0 (= vertical stress/vertical strain), is taken as
10 MPa and it is supposed that in each of the 3 finite element analyses the full
value of 0 is assigned to each new layer as it is placed.
By considering the settlement, 0, of a marker placed Y above the original ground
(assumed rigid) due to the weight of the material placed on top of it it can be
readily shown that if the material is linear elastic then the settlement profile is
parabolic, i.e.

o
:r
= 0 Y(H-Y) (1)

where 'Y is the fill unit weight and H its height.


Table 1 (a) shows the correct settlements according to Equation (1). Tables 1 (b)
and (c) show the finite element solutions respectively excluding and including the new
layer computed settlements. It can be seen that to obtain the correct results from
the finite element analysis it is necessary to use the layer boundary values only,
taking the top of each new layer as zero, i.e. the result at levels 2, 4 and 6m in
Table 2(b). These lie on the full line in Figure 2. If the mid-layer values of Table
2(b) are included the incorrect broken line in Figure 2 is obtained.
It is possible that failure to appreciate the above point led earlier research workers
to use unnecessarily thin finite element layers, ten typically being used for a large
dam. An argument for using so many was that they were required to ensure a
smooth settlement profile. From the above it is clear that as few as three layers will
give a smooth profile. Five is probably sufficient for most embankments. If more
were appropriate it would be for different reasons, e.g. to simulate a complex
construction sequence.
295

TABLE 1. One-dimensional linear elastic fill settlements (mm)


Vertical stiffness D=10 MPa. Unit weight 1'=20 kN/m 3
Figures for each layer are settlements due to that layer alone
x indicates spurious results (to be ignored)

Layer Layer Layer


IIi. Total Total Total
1 2 3 1 2 3 1 2 3
6 0 0 0 0 20 x 20 x
5 10 10 Ox Ox 19 x 19 x
4 0 16 16 0 16 16 12 x 16 28 x
3 6 12 18 Ox 12 12 x llx 12 23 x
2 0 8 8 16 0 8 8 16 4x 8 8 20 x
1 2 4 4 10 Ox 4 4 8x 3x 4 4 l1 x

(a) Analytic (b) F.e. excluding (c) F.e. including


solution new layer dis- new layer dis-
placements placements

,
- - (orrect profile
layer 3 6I~ o Layer boundary
values Icorrect)
x Mid layer values
4 0 lincorrect)
t
layer
t
2
: <) - - - -Incorrect profile

t
layer 1 l _x/
t O~'_'_
o 10 5 mm

Figure 2. Correct and incorrect interpretation of settlement in one-dimensional fill


(after Naylor and Mattar, 1988)

The above illustration is based on the assumption of a linear material. Fill is not, of
course, linear elastic and will in general have a stiffness which increases with
confining stress and therefore with depth below the surface. The procedure for
interpreting the finite element displacements is not affected by non-linearity.
Furthermore the settlement profile remains approximately parabolic (and also
symmetric) for quite significant non-linearity. To verify this consider a fill with a
296

stiffness, D, which varies with depth, h, according to

D = Do + all (2)

where Do and 0' are material constants. By carrying out a double integration the
following expression for the settlement 0 of a marker located Y above the base of
the fill and Z below its surface (H=Y+Z) is obtained:
"(Do
0 0'2
(DHLnDH - DyLnDy - DZLnDZ) (3)

in which 'Y Unit weight


~H 1 + O'H/Do
~ 1 + O'Y/D o
DZ + O'Z/Do

Table 2 compares the profile obtained from Equation (3) with that of the linear
elastic example (Equation (1 )). The symmetry of 0 about Y=Z=H/2 and the small
difference between the two profiles may be noted.

TABLE 2. Relative settlement profiles for linear and non-linear fill

Settlements are expressed as % of


Y(m) Linear Non-linear
their maximum values at mid height.
Parameters: H = 6m, "( = 20 kN/m 3
0,6 0 0 Linear: D = 10 MPa
1,5 55.6 57.3 Non-linear: Do 4 MPa, 0' 2
MN/m 3
2,4 88.9 89.6
3 100.0 100.0

It remains to show that for the one-dimensional case the stiffness of the new finite
element layer is of no consequence. This stiffness has no effect on the vertical stress
in the underlying material which is only affected by the weight of the new layer.
The new layer stiffness, of course, affects the displacement within the new layer. As
these are to be ignored (as has been shown above) it can be concluded that the new
layer stiffness had no effect on the settlement profile nor on the stress distribution.
(There is just one qualification here. The Poisson's ratio, or its equivalent for
non-linear models, must be the same in the new and underlying layers otherwise
there will be some effect on the horizontal stress.)

6. New Layer Stiffness Reduction

Although in one-dimensional fills the stiffness of the new finite element layer is of no
consequence this is not necessarily the case in general due to bending of the new
layers as gravity is applied to them. If a new finite element layer is given its
in-place stiffness value from the start then it will offer a greater resistance to bending
297

than in the real situation. This is because in reality bending occurs progressively as
the layer is built up. It would appear that some stiffness between zero and the
in-place stiffness would make the finite element layer equivalent to the real situation.
A stiffness reduction factor, f, is therefore introduced. It is defined as follows:

(4)

Q is the in-place modulus matrix(i) and Q r its reduced value for the new layer.
1 <f<oo.
The need for stiffness reduction was recognised by the early workers in this field.
Thus Kulhawy et al (1969) found that similar accuracy could be obtained using 7
layers and a large f in the 'standard dam' of Clough and Woodward (1967) as in a
14 layer analysis with f=l.
The question arises: is there an optimum value of f which can be applied
generally? It appears that there is, at least for fills modelled as linear elastic. The
value is f=4. This is obtained by making the flexural rigidity, EI, of the new finite
element layer the same as the equivalent EI of the same thickness of fill built up in
many layers. The derivation is quite subtle and is given in the Appendix.
The precise choice of f is not critical becoming less so as the number of layers is
increased, or as there is less bending in the fill (Le. as it tends to the
one-dimensional case). No optimum value has been determined for non-linear
material laws. For lack of a better criterion it is suggested that f=4 should be used
for all cases in the realisation that quite big variations from this value would have
little effect on the results.
In order to assess the sensitivity of settlement profiles to variations in both f and
the number of layers Naylor and Mattar (1988) made a study of a hypothetical clay
core dam with the clay core contained between rigid shoulders. The geometry was
inspired by the Beliche dam in Portugal(ii). See Figure 3. Rigid shoulders served to
exaggerate the bending effect in the layers and thereby produce a test situation more
severe than would normally occur in practice.
Figures 4 and 5 show respectively the effect of different layer thicknesses using a
near optimum f value, and the effect of different f values for a given number of
layers. Figure 6 shows the effect of different layer thicknesses on analyses of the
actual Beliche dam using a near optimum f value and the K-G model (iii).
Discounting the first 10m, which was excavation to bed rock over the core contact
area, the analyses comprised two, three and six layers.

(i) Q can represent either a linear or a non-linear matrix. In the latter case
it will typically relate small increments of stress to strain. It will be 3x3,
4x4 or 6x6 according to whether the analysis is plane strain, axi-symmetric
or 3-dimensional.
(ii) There is a paradox here as if ever there was a case of a dam that did
not have rigid shoulders it is the Beliche dam!
(iii) f values ranging from 3 to 5 were used for the K-G analyses. Some of
these were carried out before the optimum figure of 4 had been
determined.
298

dam
u/s t
I

,
I
\

,,
\
A_.p_-t-=t-f'
typical elem!nt
, smooth,f_+-+-l-

B-I;>-+-+-L

1
If-+-+-+ _ fixed boundaries

12m

Figure 3. Rigid shoulders finite element mesh (after Naylor and Mattar. 1988)

Legend /'; 3 layers


o 6 layers
o 12 layers
30 30

20 20

.
~
I
~
I
.
10 10

100 200 100 200


6 y (mml 6 y (mml
lal Linear elastic 11041 Ibl K-G model 11=31

Figure 4. Rigid shoulders study - effect of number of layers (after Naylor and
Mattar. 1988)
299

Legend o f = 1000
o f = 3 !lin ell.4IK-GI
.,.... o f =1
~~
30 0,:::,... 30

""~ ~
0" \
20
\ 20
I

100 200 100 200


6 y Imml 6 y Imml
lal Linear elashe Ibl K-G model

Figure 5. Rigid shoulders study - effect of 'f' (after Naylor and Mattar, 1988)

50~
40 ~
\ of
30
El
\
!
m
20

10
/
200 400 mm

Figure 6. Beliche dam centre line settlement - comparison of 2, 3 and 6 finite


element layers (after Naylor and Mattar, 1988)
300

7. Modelling Compaction

Compaction causes a temporary increase in vertical total stress as the roller passes.
There will be an associated increase in horizontal total stress some of which will be
retained.
Ingold (1979) studied the compaction of fills behind retaining walls. He showed
how the horizontal stresses are modified by compaction and how they vary with
depth. He proposed an idealised horizontal stress distribution in which a passive
lateral pressure applies from the surface to a depth depending on the effective weight
of the roller (a vibratory roller is typically double the effective weight of a
non-vibratory one). It is then constant with depth until the active pressure equals
this value. Below this the horizontal stress is assumed equal to the active pressure.
Ingold's work is useful in that it gives an indication of the depth to which
compaction affects the lateral stress and below which it is 'overtaken' by the self
weight induced stresses. This depth lies in the approximate range 3 to 5m for
compaction equipment in common use. For a large dam this would be less than the
thickness of a finite element layer.
Consequently, and because it would be very difficult to incorporate it properly,
compaction is not usually explicitly modelled in a finite element analysis, i.e. the
initial total stresses are assumed to be zero. Gravity is, therefore, the sole
mechanism for building up the stress. The effects of compaction are incorporated in
the material parameters. A stiffer and stronger material would represent a more
heavily compacted fill. (In the case of critical state-type elasto-plastic models the
initial size of the yield surface should reflect the amount of compaction (see Chapter
11 ).
In clay fills, such as in the central cores of embankment dams, compaction will
lock in suctions (negative pore pressures). Consequently the fill after compaction will
be in a state of compressive effective stress balanced by negative pore pressure. The
total stress, being (according to Terzaghi's principle of effective stress) the sum of the
two, will be near zero. In the case of the Carsington dam in England initial suctions
of about 130 kPa were estimated in the medium-stiff clay core.
This phenomena can be readily incorporated in a finite element analysis if the
analysis is carried out in terms of effective stress. In fact it is highly desirable that
it should be so that meaningful pore pressures and effective stresses are calculated for
the end of construction (or at stages during construction).
To do this an appropriate initial negative pore pressure (u o ) and equal compressive
direct effective stresses (o-xo '=0- 0 '=0- zo '=-u o ) are assigned to the fill elements. Uo
needs to be selected so that t6e calculated effective stresses and pore pressures are
approximately correct at the end of construction or of a construction stage. This will
mean that the shear strength of the clay will be correctly reproduced, an important
factor for non-linear material models incorporating an effective stress strength
criterion. It will not matter if the pore pressure and effective stress is not so
accurate just below the surface of the rising fill. Due to the changing saturation of
the soil in this region the real situation is complicated and a simplification has to be
made. This means that the initial suction (-u o ) must usually be set higher than its
actual value, as will now be explained.
Clay fills will invariably be partly saturated immediately after compaction. As the
fill level is raised the degree of saturation will increase as, due to the increasing
pressure, the air in the voids is adsorbed into the pore water. For fills placed wet
of optimum (as is now common practice) the clay will become essentially saturated
when the fill depth is considerably less than the final fill height. It is not
301

unreasonable, therefore, in this case to idealise the final clay fill as fully saturated.
Refer now to Figure 7. This represents a point in the core of a central clay core
dam where the clay has been placed wet of optimum. AC represents the effective
stress path followed during construction, which is assumed to be sufficiently rapid that
little consolidation occurs, i.e. it is undrained. C represents yielding of the clay and
it is likely that this will be reached some time before the dam has reached its full
height. The material will then be in a critical state (in the sense used in critical
state soil mechanics) so that the stress stays at point C as further deformation occurs.
OD is the corresponding total stress path. OA represents the initial suction locked in
by the compaction. CD represents the final (positive) pore pressure and the crossing
of AC and OD marks the change of pore pressure from negative to positive.
In the finite element analysis the clay is treated as saturated from the start and is
assumed to be undrained. To get the correct final effective stresses (i.e. point C in
Figure 8) it is necessary to specify an initial suction OB somewhat greater than the
actual suction OA. The effective stress path BC will then be followed in the
analysis. OB can be estimated given a knowledge of the undrained shear strength,
c u ' the effective stress failure envelope and the shape of the line BC. For linear
elastic soils and some non-linear models (including the K-G) BC will be vertical, i.e.
parallel to the ordinate. For critical state models it will be curved as shown and can
be approximately quantified.
This is the first place in this chapter where it has been necessary to distinguish
between total and effective stress methods. It is now, therefore, appropriate to
explain the finite element effective stress technique. This is done in the next section.

/
/ - '- Effstress yield line

--~---
c/
.--~-
o

/
1'
/ \ --- cr path (fe & actual)
/1. \
7 \
1-- Actual cr' \ - - Le. cr' path

U(actuall A B
- 0 t--
-u (le.1
o

Figure 7. Stress paths in the construction of a clay fill


302

8. Finite Element Effective Stress Techniques

These can be classified into two areas:


(1) Undrained loading situations where the changes in pore pressures from some
known initial value are load induced.
(2) Known pore pressure situations where both the initial and final pore pressures
are known or can reasonably be estimated. The loading may be drained,
undrained or partly drained.
Either or both of these may be used in the modelling of fill construction. (1) is
appropriate for clay fills constructed sufficiently quickly that little consolidation occurs.
Pore pressure, effective stress (or total stress as the sum of the two) fields are
calculated. The pore pressures may subsequently be compared with measurements.
(2) is appropriate for back analysis where the measured pore pressures are used as
input data. The Monasavu case study at the end of this chapter is an example of
this. Sometimes both may be used. Thus in a staged construction an undrained
analysis may be used to model a relatively rapid summer construction period followed
by a known pore pressure change analysis to model the change in pore pressure
during winter shut down. In this case if the analysis is not a back analysis using
measured final pore pressures some independent way of estimating these is required.
This may range from a guess (called a 'scenario') to some form of consolidation or
seepage analysis. The Carsington dam case study described at the end of this chapter
provides an example of the use of both methods.

8.1. UNDRAINED EFFECTIVE STRESS ANALYSIS

The concept of undrained loading implies that deformation occurs with negligible
movement of pore fluid relative to the soil skeleton. This can be understood by
imagining the soil to be made up of two components which share the same physical
space - a soil skeleton component and a pore fluid component. These are visualised
as continuous materials, both occupying all the space. This is consistent with the
continuum model used to describe the soil. Its real particulate nature is only implied.
If loading is undrained then the two components deform together, they undergo the
same strains. If there is drainage the two components move past each other like a
ghost passing through a wall.
Terzaghi's principle of effective stress for general stress states

Q: = Q:' + mu (5)

A prime indicates effective. Stress is expressed in vector form so that for 3D


applications Q: [O'x'O'y'O'zoTxy,Tyz,Tzx]T. u is the pore pressure and ill
[l,l,l,O,O,O]T. A!p'ropriate shortened forms apply to axi-symmetry and plane strain,
e.g. Q: = [O'X,O'y,T] and ill = [l,l,O]T for plane strain.
For incremental loading Terzaghi's equation takes the form

~Q: = ~Q:' + ~u (6)

where ~ indicates 'change in', for the moment not necessarily small.
For total stress analysis the material law must relate changes in total stress to
strain. A modulus matrix Q must be defined such that

(7)
303

The corresponding material law in terms of effective stress is

(8)

How are Q' and Q related for the undrained case?


Consider now the pore fluid component. It undergoes the same strains as the soil
skeleton, i.e. ~i.. As it is a fluid it only has bulk stiffness. Let this be defined by
the equivalent fluid bulk modulus Kf which relates the pore pressure change (or excess
pore pressure) ~u to volumetric strain ~fV' i.e.

(9)

To get Equation (9) in matrix form corresponding to (7) and (8) pre-multiply by
ill to give

(10)

where Q.r == mmT Kf


Since the strain vector components are mutually independent it follows from
Equations (7), (8) and (10) that

1;) ~ 1;)' + 1;)f ( II )

This is the key to effective stress finite element analysis of undrained problems. It
allows a total stress stiffness matrix to be obtained simply by adding the stiffness
contributions from the soil skeleton and pore fluid components.
The equivalent pore fluid bulk modulus Kf is related to the actual bulk moduli of
the pore fluid Kw and the soil particles Kp (not to be confused with the skeleton
modulus K') and the porosity, n, according to

n I-n
+ (12)

K p ' being the modulus of the rock minerals of which the particles are composed, will
generally be very large so that the second term on the right side of (12) will
probably be negligible.
If the soil is saturated with water then Kw is approximately 20Pa, and Kf will be
of this order. Since this is some two orders of magnitudes higher than the bulk
stiffness of the soil skeleton, K', the results of the analysis will be very insensitive to
its precise value. For this reason Kf == 2 OPa is generally a suitable value for the
undrained analysis of saturated soil.
The same formulation can also be used for drained conditions simply by setting
Kf==O. This is a nice feature of the effective stress method as it alIows the boundary
assumptions of Undrained and Drained to be investigated using the same soil skeleton
parameters D'. Only Kf is changed.
Kf values intermediate between zero and the large value suitable for a saturated soil
304

may be used to represent (1) a partly saturated soil under undrained loading, or (2) a
partly drained condition. Both are unsatisfactory. It is difficult to choose an
appropriate value for (1) due to the problem of changing saturation as air goes into
or comes out of solution. (2) is even worse. For partly drained conditions one
should either carry o.ut a formal coupled consolidation analysis, which is complex and
costly, or guestimate(l) where the result lies between the undrained and drained extremes.
The implementation of the above theory into a finite element code is
straightforward. It involves the following:
(1) Alter the data input to read separately the components of Q' (e.g. just E' and
v' if linear elastic) and Kf.
(2) Apply Equation (11) in the stiffness formulation to obtain Q so that the stiffness
matrix is identical to that in a total stress analysis.
(3) Make no changes to the stiffness assembly and equation solving parts of the
program. This means that nodal displacements and strains (Ll.) are precisely as
in a total stress program.
(4) At the stress calculation stage arrange for the separate calculation of LlQ:' by
Equation (8) and Llu by (9).
There are no restrictions on Q other than those which apply in any case to a
displacement finite element formulation. (In .. particular that it should not be singular.
This implies that Kf should not be too large(II).)
It is, in principle, possible to analyse an undrained problem either in terms of
effective stress or total stress and obtain the same results. The difference will only
be that the former decomposes the stress into its effective and pore pressure
components. Equation (11) must be satisfied. This presents a problem in non-linear
problems where the stiffness is stress dependent so that Q = QJ) and Q' = D'J').
Then there would be practical difficulties in satisfying Equations (\\ ) for varying
stress. This difficulty does not arise with linear elasticity, which will now be used to
illustrate this point.
Plane strain conditions and isotropic elasticity are assumed. Equation (\\) then
takes the form

K+G K-G o K'+G' K-G' Kf Kf


K-G K+G o K-G' K+G' Kf Kf
o o G o o o o

in which K=K+G/3, K is the bulk modulus and G the shear modulus.


It follows from Equation 12 that

}
G ~ G'
and K= K' + Kf (14 )
also K ~ K'+ Kf

(i) Technical jargon which means that an 'estimate' is really a 'guess'.

(ii) Particular care should be taken in layered analyses where ghost elements
are used with a very small stiffness together with elements containing a
large Kf. Numerical ill conditioning then becomes a real danger.
305

Example:

Suppose a saturated soil has stiffness defined in terms of effective stress by E' = 30
MPa and v' = 0.25 and an undrained analysis is to be carried out. What are the
parameters for equivalent effective and total stress analyses?
The effective stress parameters are E' and v' as above plus a suitable value of Kf.
2GPa (= 2000 MPa) is assumed. Using the standard relations between elastic
parameters

E' 30
G G' 12 MPa
2(1+v') 2xl.25

E' 30
and K' = 20 MPa
3(1-2v' ) 3xO.5

By Equation (13)

K = K' + Kf = 20 + 2000 = 2020 MPa

E and v (in terms of total stress) are now obtained from further standard relations as

9KG 9x2020x12
E 35.9 MPa
3K+G 3x2020 + 12

3K-2G 3x2020 - 2x12


and v 6x2020 + 2x12 = 0.497
6K+2G

These, therefore, are the total stress parameters which correspond to E', v' and Kf.
Note that if Kf were made very large v would tend to 0.5 and Q. would become
singular.

8.2 KNOWN PORE PRESSURE CHANGE ANALYSIS

The basis of this technique is that the increase in pore pressure (L1u) in effect applies
a body force of intensity Q to the soil skeleton according to:

Q = -'i7(L1u) (15)

'i7 is the gradient operator so that Q is the vector force/unit volume.


In the finite element analysis the nodal forces corresponding to Q are calculated in
the same way as they would be for any other body force (such as gravity) and added
to any other applied loads.
The appropriate stiffness for this analysis is clearly that of the soil skeleton, i.e.
Q.'. It is in the category of a drained finite element analysis (Kf=O). It need not,
however, be physically 'drained' (which implies dissipation of excess pore pressures)
but simply that the pore pressure change is known. Thus it could apply to a partly
drained situation.
Figure 8 illustrates the technique applied to a water table rising 3m in a sand.
306

Here Q is simply Archimedean uplift on the newly flooded sand, i.e. 10 kN/m 30).
The net uplift would actually be less than this due to the added weight of the water
filling the voids of the 3m sand layer. It would probably be 7 or 8 kN/m 3

. ,
y

~I
In,tiol p =-~ lJu
Y ay
o
I.. 30
"'1 KPo 10 KN/m 3
Pore pressure Body force

Figure 8. Known pore pressure change technique applied to rising water table (after
Naylor, el ai, 1983)

9. First Filling and Operation - General

What follows mainly applies to embankment dams, particularly those with internal
cores. Some of it, however, is also relevant to fills which are not dams where for
some reason a water tables rises. The sections dealing with the water loading and
collapse settlement apply here. A much simpler one-dimensional approach which
need not involve finite element analysis may, however, be sufficient for such fills.
The stages of interest following the construction of an embankment dam are first
the end of impounding, then the 'steady seepage' condition which will be attained in
pervious core dams some time after the reservoir has been filled, and, finally, rapid
drawdown.
Compared with the modelling of construction relatively little has been done in
applying the finite element method to these subsequent stages. Here attention is
focussed on the impounding. Brief treatment is also given of the steady seepage
condition but as will be seen this can be analysed as a small variation on the case of
slow impounding. Finite element techniques for modelling drawdown are not covered.
Little work has been done on this.
The post construction performance of dams containing a zoned internal core are
complicated. This is illustrated in Figure 9 which shows in a qualitative way how the
pore pressure at the point X could be expected to vary during construction and
reservoir filling. Prediction of the pore pressure is diffult because of the simultaneous
actions of the dissipation of excess pore pressures caused by physical movements of

(i) The unit weight of water is taken here, as elsewhere, as 10 rather than 9.81
kN/m 3
307

the dam (due to the weight of the dam and the reservoir) and the pore pressures due
to the penetration of the reservoir water into the core.

Pore
pressure
at X

I.constructiofll Impounding 1_ Res. full


Hypothetical dam
(3 years)

0. iii 1'111: Construction*


Cause of
pore pressure { '::.>. Impounding (a) Reservoir load*
(b) Collapse settlement-

Seepage
Excess pore pressure due to Imposed loads on core

Figure 9. Pore pressure changes in central clay core dam

A rigorous modelling which would take into account this 'coupled flow' problem
requires the use of a coupled consolidation analysis incorporating the Biot theory.
This is difficult both from the point of view of the analysis itself and because of the
difficulty in assuming meaningful permeabilities. Naylor, Knight and Ding (1988) did
manage to apply the Biot theory in the back analysis of the Monasavu dam and
obtain pore pressures in reasonable agreement with the piezometer measurements.
However, as this was a back analysis, it was possible to adjust the permeabilities to
obtain the agreement. This cannot be done in a predictive analysis. This method is
not covered in this chapter and the interested reader is referred to the above paper
where the theory is also given.
For dams where the internal core pore pressures are important the approach taken
here as an alternative to a coupled consolidation analysis involves either bounding
assumptions as to whether conditions in the core can be assumed to be 'drained' or
'undrained', or the assumption of pore pressure 'scenarios' . This is the
known-pore-pressure-change technique of the previous chapter. The pore pressures
need to be estimated separately from the finite element analysis.
A most important factor affecting internal core dams is that of col/apse settlement.
This is the phenomena whereby the saturation of rockfills causes them to settle (i). It
is most marked in poor quality rockfills particularly if they are not heavily compacted.

(i) The writer has also called this 'saturation shrinkage' as being descriptive without
the connotation of collapse in the sense of a complete failure of the dam.
'Collapse settlement' has, however, wide usage; but the title should be used in
full, as to talk of 'collapse' implies failure!
308

The Beliche dam in south-east Portugal exhibited significant collapse settlement. This
was an intended design feature. It is described in the case studies section at the end
of this chapter. The EI Infiernillo dam in Mexico (Marsal and Ramirez 1967)
provides a classic example. This 148m high dam has a narrow central clay core.
During reservoir filling the core first deformed downstream (as would be
expected) but by the time the reservoir had reached its full height collapse settlement
in the upstream shoulder had pulled the top of the core back until it was
approximately where it was at the start of impounding. This effect forms an
important part of the finite element analyses of impounding.
The water loading on and within the dam and foundation (assuming this is
compressible and/or pervious) is considered next. Then in Section 11 the finite
element procedures for modelling the impounding and the development of steady
seepage are covered. Although needed for the analyses of Section 11 the techniques
for modelling collapse settlement are dealt with after it in Section 12.

10. Loading due to Impounding

10.1 UPSTREAM MEMBRANE DAM

This is the simplest case. A hydrostatic pressure due to the water load acts on the
upstream face (Figure 10).
If the foundation is compressible and is included in the finite element mesh then
different idealisations of the loading are appropriate according to the physical
situations. The following will cover most cases:
(1) Impervious foundation (Figure 10a)
(2) Pervious foundation initially dry with an impervious vertical grout curtain
below the toe of the membrane (Figure lOb)
(3) Pervious foundation with initial water table at the ground surface and a
vertical grout curtain as for (2) (Figure 10c).
The loadings illustrated in Figure 10 correspond to what is here referred to as the
external load method. This will certainly be appropriate for the loading on the
upstream membrane. However, for foundation loadings the alternative internal loading
method described below may be more convenient.

10.2 INTERNAL MEMBRANE DAM

This type of dam is not very common nowadays, but as the loading is similar yet
simpler than the general case of a pervious zoned dam of fill it serves here to
introduce the external and internal load methods. They are alternatives. Both
represent the same physical situation.
The loading differs from that of the upstream membrane dam only in that an
Archimedean uplift acts upstream of the membrane. This is shown by the upward
pointing arrows in Figure 11.
Such a loading would cause the upstream part to rise up. This does not happen
in practice as the Archimedean uplift is more than compensated by collapse
settlement.
309

External load method


Generally soil can be considered as a two component material compnsmg soil skeleton
and pore fluid parts. In the external load method only the soil skeleton is
represented by the finite elements. The pore fluid exerts an external load on it.
Thus when the upstream part of the fill is flooded (Figure 11) an upwards acting
body force (force/volume) is applied to the submerged elements.
This may be viewed as a total stress method because it does not explicitly take
account of pore pressures. The pretence is made that the flooded soil is dry! The
final stresses computed in the analysis are the effective stresses. The analyst must,
himself, add the pore pressures if he wants to know the total stresses.
The regions not permeated by the reservoir water may be analysed in terms of
either total or effective stress. If, as is usually the case, they comprise either dry
rockfill or free draining material the distinction becomes academic since the two

~Foundahon elements
(If required) -------

(a) ImpervIous foundation

t t___
(\'"
-I Assumed ImpervIous grout (urtam
t t t-~
(b) Pervious foundation initially dry

Initial and final water table


/ ' dis of grout curtain
~----~~~~~--~--

grout (urtain

(c) PervIous foundation. initial water


table at ground surface

Figure 10. Reservoir loading on external membrane dam


310

..
Assumed ImpervIous
For al ternahve foundahon
I__ grout curtain
loadings see Figure 2 I

NB PA = o.,!1-n.1 where n. air porosity prior to filling

Figure 11. Reservoir loading on internal membrane dam - external load method

methods then become essentially the same(i).


The disadvantage of the external load method is that the pore pressures are not
included in the analysis. The internal load method does not suffer this disadvantage.

Internal load method


Figure 12 illustrates the loading to be applied. It is only required on the mesh
boundaries. The internal loading is obtained by prescribing the final and initial pore
pressure fields. The method is, in fact, the known-pore-pressure-change technique
described in 8.2 above. The body forces corresponding to the Archimedean uplift,
etc., are calculated within the program as the gradient of the pore pressure change.
This is a true effective stress method. Not only does it have the advantage that it
automatically includes the pore pressures in the analysis but it also allows the analysis
of more complicated situations than that of simple hydrostatic uplift. These appear in
the case of seepage through internal core dams which are considered in the next
section. It is not practical in this case to work out appropriate body forces
corresponding to the seepage forces for use in the external load method.

(i) Precisely the same for dry materials, but for free draining material below the
water table the pore pressure will be included in an effective stress analysis as a
'passenger', i.e. it will have no effect on the displacements nor on the stress
changes.
311

U o = Inherited or assumed 1 uf = Uo
u f = '6 . . h'" (If no seepage) L Impervious grout curtain

Uo = Initial pore pressure (inherited from construction


analysis)
u f = final pore pressure

Figure 12. Reservoir loading on internal membrane dam - internal load method

Uf assumed
Inherited

uo=lnherlted* Uf
u f ='6 . . h .... I varies I

us uall y ='6 .,(h .... -H .... )


~ 1
Grouted zone

Figure l3. Reservoir loading on central core dam - internal load method - slow
filling
312

10.3 ZONED EMBANKMENT DAMS

This type of dam contains a more or less central zone of relatively impervious
material flanked by shoulders of, typically, rockfill. The central zone or core is
usually clay but can be non-plastic comprising silt/sand mixtures. The shoulders are
usually relatively permeable. Their purpose is to support the core.
The difference between this type of dam and the central membrane dam considered
in 10.2 is that the time taken for the development of seepage through the core needs
to be taken into account. This is complicated as illustrated in Figure 9. Two
extreme conditions may be considered:
(1) Filling sufficiently rapid that negligible penetration of water into the core
occurs during impounding.
(2) Very slow filling so that steady seepage is established in the core by the
time the reservoir is full.
The truth, of course, will lie between these extremes. The second may be a
reasonable approximation of the situation some time (Le. some years) after filling if
the reservoir has been kept more or less full.
Condition (1) is idealised by pretending that an impervious membrane exists on the
upstream face of the core. The loading then is the same as that for an internal
membrane dam (Section 10.2, Figures 11 and 12). The external or internal load
methods may be used. The core would be idealised as undrained.
For Condition (2) an assessment of the final pore pressure field in the core is
required. This may be determined independently of the finite element analysis, e.g.
by a steady state seepage analysis. Alternatively a pore pressure 'scenario' may be
assumed. (The assumption of a number of different scenarios to represent extreme
possibilities is in the writer's opinion a very good way of using the finite element
method.) If the analysis is a back analYSis the final pore pressure field can be based
on piezometer measurements. (Not precisely because there are always vagaries in
such measurements.) The internal load method must then be used, i.e. the analysis
becomes a known-pore-pressure-change analysis in which the initial and final pore
pressure fields over the whole mesh are provided as data (Figure 13).
The reservoir filling analysis will normally be preceded by a construction analysis
which will provide the end of construction stresses, and pore pressures if it is an
effective stress analysis, as initial values for the reservoir filling. There will usually
be initial excess pore pressures in the clay core, and possibly in the foundation as
well. These may have resulted from an undrained construction analysis, or from a
construction analysis involving the known-pore-pressure-change technique to make the
end of construction pore pressures agree with measurements. In any case,
complicated pore pressure distributions in the core can be expected thus underlining
the need for the internal load method of analysis in this case. In the shoulders the
situation will be simpler. In both there will usually be no pore pressures at the end
of construction, and then on impounding there will be a hydrostatic distribution in the
upstream shoulder and probably still zero pore pressure in the downstream shoulder.

11. Analysis of First Filling and Operation

11.1 FIRST FILLING

This can involve just a single finite element analysis in which initial stresses and pore
pressures are inherited from a preceding construction analysis and the loadings are as
313

given in the preceding section. (Displacements and strains can, if desired, also be
inherited.) If collapse settlement is a factor, as it will be in most dams except the
upstream membrane type, then obviously the finite element program must incorporate
the theory described in Section 12 below.
For major dams it will be desirable to carry out a staged analysis. This will not
make much difference if there is little or no collapse settlement (and none at all if
the material is modelled as linear). Where significant collapse settlement occurs then
the number of stages should be similar to the number of layers used for construction,
perhaps in the range 3 to 6.
The procedure for a staged filling analysis is quite straightforward and free of the
subtleties of displacement interpretation and new layer stiffness reduction which occur
in construction analysis. Figure 14 illustrates the loading required using the internal
load method for the central core dam case. Note that the pressure distributions for
each stage are not the total-to-the-end-of-that-stage but are incremental, i.e. a
linear pressure increase over the height of the current stage and constant below that.

rI p;r:::r.::I~~_"':U~O":':=':'O:':'U:!.f":':=l\~wh,
uo=l\.,,(h,-H,)
(a) Stage 1
Uf=l\w h,

Uo=l\w h,
Uf=l\whZ

UO="l\.,h2 (e) Stage 3


u f ="l\.,h 3

Figure 14. Three stage reservoir loading for dam of Figure 13


314

11.2 STEADY SEEPAGE CONDITION

This implies an internal core dam. A known-pore-pressure-change analysis in which


the initial pore pressures are those at the end of reservoir filling, and the final
obtained from a separate steady seepage analysis, is required.
Two boundary assumptions may be made. The first is that the reservoir is filled
rapidly and the second that it is filled slowly. In the first case an impounding
analysis must be carried out first followed by the development-of-steady-seepage
analysis, i.e Case (2) of the impounding analyses described in Section 10.3 above.

11.3 FINITE ELEMENT CONSIDERATIONS

Here the detail of how the loads of Section 2 can be applied is considered. This
concerns the input provisions in the finite element code.
The loadings involve the application of
(1) Surface tractions acting normal to element sides (faces in 3D) and varying
in intensity across the side.
(2) Body forces.
These loadings must at some stage be converted to nodal forces to form the finite
element load vector. This can either be done by the analyst or done within the
program. The latter is virtually essential and all finite element codes intended for
geotechnical use should have this facility. The theory for deriving the nodal forces
corresponding to surface tractions and body forces is not given here. It is based on
the virtual work method and the reader is referred to standard finite element texts.
It is covered in Chapters 2 and 3 of Naylor et al (1983).
Not all codes, however, provide for variation of surface pressures and body forces
across elements. If this is not possible then a stepped distribution of pressure would
be required with consequent error. One way of inputting a smooth variation of
tractions along element sides is to assign sets of tractions (in this case the pressure
normal to the element surface) to element nodes. The element side has also to be
identified. Thus in Figure 4 if there were five layers of elements over the depth HU)
six surface pressure sets ranging from zero at the surface to 'YJiw at the bottom
would be input.
A similar facility may also be used for prescribing varying body forces. This,
however, is not normally necessary as a single value can usually be applied to a
complete element. (It is noted in passing that this procedure for assigning values by
sets to nodes and allowing the program to calculate their distribution over the
elements can also be used for material properties, initial stresses, etc. This is highly
desirable in geotechnical applications where significant variations across elements
commonly occur.)

12. Collapse Settlement

Failure to incorporate this important phenomena in a reservoir filling analysis will


result in an internal core dam deforming upwards in its upstream regions, a
phenomena which has probably never been observed!
A theory for doing this was developed by Nobari and Duncan (1972). It is very
much tied to the use of the 'hyperbolic' stress-strain law of Duncan and Chang
(1970) and relies on the triaxial test for its parameters. As it underlies much of the
work done to date in this area it is outlined here. The writer more recently has
315

been working on a generalisation of Nobari and Duncan's method so that it is not


tied to a particular stress-strain law. This is described by Naylor et al (1989) and is
presented in 12.2 below.
The collapse settlement theories are based on the following experimental finding.
Two identical samples of initially non-saturated soil are tested. One sample is
saturated and then loaded. The other is first loaded and then saturated. It has been
found that the final stress and strain states are nearly the same. This is illustrated in
Figures 15 and 16.
This finding suggests that two sets of material properties may be used to describe
the soil - one the soil as placed, i.e. before saturation, the other the saturated soil.
The stress-strain curve of the dry soil is followed until saturation occurs. Then a
switch is made to the saturated soil stress-strain curve. This is done in two parts,
first a stress change without strain and then a relaxation of the resulting
out-of-balance stress. It is followed by the application of the impounding forces
described in Section 2 above. Thus the loading has to be applied in two stages:
(1) Collapse settlement loading involving
(a) determination of saturation stress change under no strain,
(b) determination of nodal forces to restore equiliorium.
(2) Reservoir loading.
The above applies to both Nobari and Duncan's method and the writer's
generalisation of it. The main difference between the two methods is in the detail of
1 (a).

0.80

0
0.75
+=
IV
0:
'C
0
>
0.70

0.65

0.1 10 10 30
AppLied Load-tons per sq ft

Figure 15. Effect of saturation in oedometer tests on silt (after Burland, 1965)
316

15
s
\
\
10 I
\ (3)
S \ ---
- -.. . . , V ....
\ _ - - - - - (1)
........

o
o 5 10 15

- -_ _ _ _ _ 5

" "-, "-


' ....... (3)
5
S Indicates saturation pOint
(1),(3) cell pressure kg/cm2

Figure 16. Triaxial tests on soft rockfill with saturation


317

12.1 NOBARI AND DUNCAN'S METHOD

In order to determine the stress change under conditions of no strain due to


saturation Nobari and Duncan carried out consolidated drained triaxial tests on the dry
and saturated materials. These would be repeated for different ceJI pressures to cover
the range of interest in the dam.

Dry sOil ~ III <E-


~ ; ~p
Saturated sOil - - - - ~
...
Ill"
~
~

(d)
O"d
0"3=P O"d
(5)
S/,-0"3=P
.A'
/1
/ I
/
0
Ea 0 / Ea
A B
0 Ea 0 Ea

b \B b
", ........

Ey
a D E y\
a 6'-
(a) Conventional presentation

o B A o ___ 060~ _ _ _ _; ' Ev


BA
k-p(d)~1
(b) Alternative presentation

Figure 17. Nobari and Duncan's method for clamped saturation stress change
318

Stress space
0
(bl
D,S

~n
These figures illustrate the
nonlinear case. For linear
elastic materials all the lines
space
00 and as are straight a
(cl

o = Stress,strain origin

Figure 18. General method for clamped saturation stress change (after Naylor et ai,
1989)

Figure 17(a) illustrates the procedure. First a test on the dry soil is carried out at
a selected cell pressure p(d). It is required to find the saturated deviator stress (T d(s)
which corresponds to the dry deviator stress (Td(d) indicated by the point D, and also
the saturated cell pressure (or minor principal stress) pes). In the left hand diagram
OA represents the consolidation stage and AD the drained shear stage. The point D
on the axial strain:volumetric strain plot is then identified. It is given the coordinates
a and b. The crux of the method is to carry out a test on the saturated soil which
will produce these same strains, i.e. with an fa: fV plot which passes through the
point D.
The detail of this is not given here, and the reader is referred to both the original
paper and Maranha has Neves and Veiga Pinto (1988) for a fuller explanation.
Suffice it to note that triaxial tests on the saturated material are carried out so as to
obtain the stress-strain plots in the right hand diagram of Figure 9(a). OB represents
the consolidation stage to a cell pressure pes). Point S on the (Td:fa curve then
defines the required saturation stress, (TiS).
The required clamped stress change is therefore from (Td(d),p(d) to (Td(s),P(s). This
provides the basis for part (b) of Stage 1, i.e. the determination of nodal forces to
effect the release of the imaginary clamps. The detail of this is very much tied to
the formulation of the hyperbolic model of Duncan and Chang. (See again Maranha
das Neves and Veiga Pinto, 1988.)
Figure 17(b) provides an alternative presentation of the Nobari-Duncan procedure
which allows it to be compared more easily with the method described in the next
section. (Tm is the mean stress which for the triaxial test is (P+(Td/3), and fd is a
deviatoric stress measure = (fa-fyf3). Other symbols are the same as in Figure
17(a).
319

12.2 GENERALISATION OF NOBARI AND DUNCAN'S METHOD

The procedure for determining the 'clamped' stress change of Stage 1 (a) is first given.
This is the only part which differs from the Nobari-Duncan method. Next, Stage
l(b) is outlined and this applies to both methods. Stage 2, the application of the
water load, is not described here as this has already been covered in Section 10
above.
No reference is made here to the soil tests required to obtain the dry and
saturated material properties. They do not affect the theory. It is supposed that
suitable tests - triaxial, oedometer, or whatever - are carried out to define a dry
modulus matrix Q(d) and a saturated modulus matrix Q(s). These matrices will
correspond to the particular material law in use and will, in general, be stress
dependent and perhaps strain dependent as well.
To calculate the stress change for a newly saturated point in the fill it is supposed
that the dry soil is loaded from a state of zero effective stress to its pre-saturation
(calculated) stress state Q:(d). A straight stress path (OD in Figure 18) is assumed.
The strain path corresponding to the stress path, OD, is calculated. Note that in
general it is not sufficient just to calculate the strain, .s, corresponding to OD as the
path in strain space will not necessarily be straight. The saturated soil is then
supposed to be loaded up this strain path until the strain, .s, is reached. The
resulting stress Q:(s) is the required clamped post-saturation stress state.
This may be expressed as follows. Let the path OD be divided into n increments,
Ll(Ti(d). n will be chosen to suit the non-linearity of the stress-strain law. n can ::
1 if the material is linear elastic. Let Llf indicate the corresponding strain increment.
Then for an increment i

(16)

(17)

Here 12.(d) and 12.(s) are respectively the dry and saturated modulus matrices in terms
of effective stress. The clamped saturation stress is then

n
(T.(s)
-1
2 LlQ:i (s) (18)
i=l

In the program the path 00 will first be followed and the Ll.si stored in an array.
The strain path will then be followed and Q:/s) accumulated. This is done for each
'integrating point -(i).
This completes Stage 1 (a). The corrective nodal forces Q required to restore
equilibrium are then calculated from

(19)

(i) For numerically integrated finite elements. These will usually be Gauss points.
320

Here !! is the standard geometric matrix used in finite elements to relate strains to
element nodal displacements, Qe is a vector of element nodal forces and the
integration is taken over the element volume.
The Qe are assembled for each node (i.e. the contributions from elemp-nts sharing
the node added) to give the global nodal force vector Q. The solving of the finite
element equations for this load vector comprises Stage l(b).
Note that the stress path OD in Stage 1 (a) and the corresponding strain path are
not actually followed by the soil. They are simply a theory to enable (J(s) to be
calculated. The experimental findings (Figures 15 and 16) do suggest that this is a
reasonable theory. The selection of a straight line for OD is quite arbitrary,
however, and work needs to be done to determine the effect of different assumed
shapes for OD. The shape of this line identifies a difference between this and the
original Nobari-Duncan method. The line OD for the latter is kinked, i.e. it is the
line OAD in Figure 17(b).
For isotropic linear elasticity an explicit expression relating the clamped saturated
stress components to the dry stress can be obtained. This is readily derived by
introducing the components of the elastic Q matrices into Equations (16) and 17).
For plane strain, and using the 'K,G' form of the D matrix of Equation (13), one
obtains:

Q:( s)
[
(Jx
(Jy
T
(, [ rK+rC,

Symmetric
rK-rC,
rK+rC.
0
0
2rC
1[
(Jx
(Jy
T
(' (20)

K(s) C(s)
in which rK= --=-( d) , (K=K+C/3) , and rC - c(d)
K

The procedure used by Professor Justo and described in Chapter 6 fits into this
formulation as a special case. He assumes that Poisson's ratio is unchanged by
saturation so that rK=rG=(I-a) where 'a' is determined from tests. Substituting this
in (20) results in

Q:(s) = (l-a)Q:(d) (21)

Note that Equation (21) is only valid for the elastic case in which the dry and
saturated Poisson's ratios are the same.

12.3 ONE-DIMENSIONAL EXAMPLE

To clarify the foregoing a simple example comprising the saturation of a laterally


constrained column of elastic soil 6m high is worked out explicitly. The column is
not divided into finite elements as in this case the method can be more easily
demonstrated by reference to the continuum solution. The material properties are as
follows. D is now the scalar vertical stiffness.
For loading D(d) = 9 MPa, D(s) = 6 MPa
For unloading D is increased 4 times
Bulk unit wt., -y, = 18 kN/m 3
Archimedean uplift, PA = 10 kN/m 3 (u = -YO)' for simplicity
321

the air voids existing before flooding are neglected)


Initial pore pressure zero =
Setting n=1 (since linear elastic) and dropping the ,1 prefix, Equation (16) becomes

f =

Now <red) = "Y.(H-y) = 18(6-y) kPa in which y is the height above the base.
Substituting D(d) = 9000 kPa:

18(6-y)
E = 9000

Introducing Equation (17)

<res) = 6000 ( = 12(6-y) kPa

and <red) - <res) = 6(6-y) kPa

This out-of-balance stress is equivalent to a body force:

q = 4- (<r(d) _ <r(5 = 6 kN/m 3 (22)


dy
acting upwards. (22) corresponds to Equation (19) except that (19) computes the
nodal forces for the finite element calculation.
Application of q downwards will simulate the removal of the imaginary clamps and
will restore equilibrium. The settlement associated with this is obtained by noting first
that the strain caused at level y is

<r(d) - <r(5) 6(6 - y)


f = 0(5)
Y 6000
The settlement at level y=y is then

Y
(12 - Y)Y
c5y( 1)
J fydy =
2000
m (23)
0
Stage 1 is now complete. Figure 19 shows the stress :strain path for Stages 1 (a)
and 1(b) followed by a point at mid-height. The settlement profile corresponding to
Equation (23) is shown as the line (1).
Stage 2 comprises the upward application of an Archimedean body force of 10
kN/m 3. If the soil stiffness remained equal to D(s) this would add -10 by(l )/q to
the Stage 1 settlement. In this example, however, the soil is four times stiffer on
unloading so that the movement would be a factor of 4 less, i.e. <ry(2) = -2.5
(Ty(1 )/q with q = 6 MPa. Adding this to Equation (23) gives

7(12 - Y)Y
m (8)
24000
322

This is shown as line (2) in Figure 19. The associated stress:strain line is labelled 2.
Stage 2 is now complete.

By(mm)
0 10 20
6 60
0

Y(m)
4

2
kPa
C1
40

20
1/ /'
Oldl~
.
/,"
/1a
1(/
. IS
I
2
J
1
I,.Olsl

~ ~Olll
/1 I
0
o
Fictitous 0.5 Actual
Strams ( % ) -
By at end of
stages (1) & (2) Effective stress path at Y = 3m

Figure 19. Collapse settlement and uplift for one-dimensional example

Note the stiffness assumptions made in this example. The post-saturation stiffness
is assumed for Stages l(b) and 2. Also note that the soil is considered to be subject
to loading in Stage 1 b but to unloading in Stage 2.
323

13. Applications

Three dams on which finite element back analyses have been carried out are
described here. They comprise the Carsington dam in the U.K., the Beliche dam in
Portugal, and the Monasavu dam in Fiji. All three illustrate the procedures for
modelling construction, and the last two impounding as well. All three involve
known-pore-pressure-change analyses. Beliche is of particular interest for its collapse
settlement modelling. It is the only one of the three in which this was modelled.
Carsington is of interest because the dam collapsed at the end of construction. The
description given here of these three case studies is a slight modification of that given
by Naylor (1990).

13.1 CARSINGTON DAM

This unusual dam was constructed entirely of clay between 1982 and 1984 and
collapsed as construction was nearing completion in early June 1984.
The finite element work carried out by the writer and described here comprised a
back analysis of the dam. Its objective was to determine the mechanism of failure.
The emphasis was on a realistic modelling of the stages of construction to determine
the development of stresses, pore pressures and deformations. Parallel finite element
work was carried out at Imperial College, London and is reported Dounias et al
(1988), and Potts et al (1990). The failure itself is described by Skempton and
Coates (1985) and then in detail in Coxon's (1986) report on it. Figure 20 shows a
cross-section on the valley centre. Here the failure slip surface passed through the
core, its 'boot' and then close to the base of the dam (in places passing into the
foundation) to emerge near the upstream toe.
The analyses were carried out in terms of effective stress using the K-G model
described in Chapter 11.
The summer construction seasons were idealised as undrained in the core but with
full drainage assumed over the bulk of the shells. (An intermediate assumption was
made for a transition zone around the core.) This idealisation allowed a non time
dependent effective stress analysis to be carried out.
The use of a high value of the equivalent pore fluid bulk modulus, Kf, in the core
idealised this material as a saturated clay. It was given the necessary shear strength
(c u in the region of 50 to 60 kPa) by the assigning of an initial suction of 130 kPa.
This simulated the effect of compaction as has been explained in Section 7 above.
Kf was set to zero in the shells (except in the transition zone where an intermediate
value was chosen).
During the winter period between the end of one construction season and the start
of the next the pore pressures adjusted as consolidation took place. The dam was
well supplied with piezometers (14 in the core at the mid-valley section) and these
provided a good record of the pore pressures. Advantage was taken of this to carry
out known-pore-pressure-change analyses to model the winter shut down period. An
approximation of the winter-end pore pressure field was input as the final pore
pressures. The initial pore pressures were inherited from the preceding undrained
construction analyses. These would not necessarily be in agreement with the measured
end of construction season values. (The extent to which they differe~ is or
considerable significance as will shortly be explained.)
In all six sequential analyses were used to model construction: OD, 1 U, ID, 2U,
2D, 3U. The number indicates the construction seasons, i.e. 1 :1982, 2 :1983, 3 :1984.
o is a preliminary analysis to model the foundation excavation. U and D indicate
324

Undrained and Drained respectively. Undrained being the construction stage and
Drained the known-pore-pressure-change stage. Figure 20 shows the main zones and
the levels reached at the end of each season.
Figure 21 compares measured and calculated pore pressures at the end of the 1983
construction season (when the bulk of the fill was placed) and indicates by shading
regions where the difference between measured and calculated is significant (> 50
kPa).
These differences illustrate the role of the finite element method as an arbiter
against which to compare measurements. Something unusual is happening in the
shaded regions, suggesting a need for further investigation. At Carsington the cause
of the unusually high pore pressures still awaits a full explanation. It was clearly an
important factor in the failure.
Another area of interest in the Carsington work was the modelling of the failure
mechanism. The analyses showed the development of a sheared zone which started
inside the dam (in the toe of the boot-shaped core) and spread outwards following
approximately the recorded slip plane. This aspect of the work, namely the use of
the finite element method to predict a shear failure, is outside the present scope. It
is discussed briefly by Naylor (1990)

:.;..,;;...;...,m~...:::----.L 19 84 'V: E1.19 7


--~
V' Et.182

Foundation Base of f.e. mesh

NOT E Fill up to El.197m was discretized into 5 layers of 'parabolic' elements.


Top 3m were excluded from mesh. A surcharge was applied in lieu.

Figure 20. Carsington dam valley centre section (after Naylor, 1990)
325

Measured pore pressure> SOkPa above calculated


----_ .., Measured pore pressure> SOkPa below calculated

Figure 21. Carsington - comparison of end of 1983 construction pore pressures with
measured values (after Naylor, 1990)

13.2 BELICHE DAM

This SSm high dam constructed between 1982 and 1985 has been the subject of
collaborative research between the Department of Civil Engineering in Swansea and
LNEC(i) in Lisbon, Portugal for a number of years. This started with separate
predictive analyses of the construction of the dam which were carried out before the
dam was built. The predictions are described by Naylor et al (1986). Subsequent
work has concentrated on the modelling of the reservoir filling and the performance
since then. The main points from both the predictions and the subsequent work are
outlined here.
The predictive analyses of the construction stage are of interest firstly in that they
compare predictions carried out independently using both different material models and
different finite elements, and secondly because they provide results for comparison
with the actual performance.
Figure 22 shows the meshes used for the LNEC and Swansea analyses and Figure
23 the settlement profiles on three vertical sections for the end of construction
predictions. It can be seen that the maximum settlement is near the middle of the
dam and is about 40cm. There is not a great difference between the results from
the two institutions.

(i) Laborat6rio Nacional de Engenharia Civil


326

Unfortunately it is not possible to add a third 'as measured' line to Figure 23


which would be meaningful. This is because unusually heavy rains caused partial
impounding and consequent saturation of a major part of the upstream shell before
construction was complete. Thus collapse settlements occurred which were not taken
into account in the predictive analyses. The maximum settlement in the middle of
the dam exceeded 1 m.

~ CENTRAL CORE

_"'LTEII
c::J INNER SHEll

~ OuTlER SHELl
~ ALLUVIAL FOUNOArlON

(a) L NEe finite element mesh

11 13 IS - Inclinometer locations

(b) Swansea finite elemen~ mesh

Figure 22. Beliche dam - finite element meshes (after Naylor et at, 1986)

Back analyses based on the post- -construction settlement records show that the
original predictive analysis underestimated the stiffness slightly. Had premature
impounding not occurred, i.e. if constuction had been as assumed in the analyses, the
maximum settlement would have been about 60cm, i.e. some 50% higher than
predicted. The error is probably mainly accounted for by differences between the
pre-construction test samples from which the material parameters were obtained and
the actual fill material. Other factors are the numerical models themselves and
327

possibly some scale effect although this factor was minimised by the use of large scale
tests. The error due to the finite element discretisation would be minimal.
The analyses carried out after impounding when the deformation measurements were
available took into account the collapse settlement using the technique described in
Section 12.2 above. The analyses were quite complicated as they had to take into
account the collapse settlement in the upstream shell which occurred before the dam
had been built to its full height. The computed and measured settlements are
compared along settlement markers 11 and I3 in Figures 24 and 25 respectively. The
big step between the broken lines 3 and 4 in Figure 24 shows the contribution of the
collapse settlement to the total. No fill was added between these stages. It can be
seen that by taking collapse settlement into account quite good agreement with
the-end-of-impounding settlements could be obtained. This, however, does not take
into account the settlements since then which are due presumably to further
rheological effects. This is shown by the movements between MarchI April 1985 and
9th January, 1989.

- + - L.N.H
-0-- Swansea

-10 10 20 em 10 20 30 em 10 20 30 40 em
Horizontal Vertical Vertical
11 13

See Figure 2l for locations

Figure 23. Beliche dam - predicted deformation profiles on Sections 11 and I3


32R

55,----,-----,----,-----,-----~--~

so 11th Apr. '85

45 -"

40
"'\ \31 '85
st Jan.

\-<D
\
E

25

20

15

10
- Measured
- - - K-G model
5
(S)-- Analysis stage
O~--~----~-----L----------------~
o 200 400 600 800 1000 1200
Settlement (mm)

Figure 24. Beliche dam - measured and back-analysed settlement profiles on II


329

55

50

45

40

35

5. 30
+-
.c.
CTI
iii 25
I

20

15

10

5 (Legend as for Fig.24

0
0 200 400 600 800 1000 1200 1400
Settlment (mm)

Figure 25. Beliche dam - measured and back-analysed settlement profiles on 12


330

13.3 MONASAVU DAM

This 85m high rockfill dam comprises a soft clay central core between hard rockfill
shells. It was constructed between 1979 and 1982. It is unusual for such a high
dam to have so soft a core as placed (c u after placement averaging 17 kPa). This is
discussed by Knight el al (1982). It is comprehensively instrumented and has been
back analysed by the finite element method. These analyses are described by Knight
et al (1985) and Naylor et al (1988). The former of these provides an example of
probably the first published use of the known-pore-pressure-change technique. It also
models the fill as linear elastic. Highlights from the latter paper are presented here.
The 1988 paper differs from the 1985 one in two respects: a coupled consolidation
formulation incorporating the Biot theory was used thus making the calculation of pore
pressures deterministic, and an elasto-plastic soil model for the soft clay core was
introduced for comparison with the elastic. The elasto-plastic model was the
'continuous plasticity critical state model' (c.p.c.s.m.) described by Naylor (1985).
The analyses covered construction, impounding and one year of operation.
Comprehensive deformation, pore pressure and stress (horizontal and vertical)
measurements were available in the core at three elevations throughout this period.
The analyses were not therefore predictive except in so far as they extended the
measured data to other parts of the dam. The parameters (which included the
permeability of the core) were adjusted to give as good a fit as possible to the
measurements.
The finite element mesh is shown in Figure 26. Figure 27 compares the elastic
and c.p.c.s.m. computed settlements and piezometric elevations with the measurements.
Figures 28 and 29 compare the computed and measured vertical total stress and pore
pressures on a horizontal section near the base of the dam at, respectively, the end
of construction and about one year after reservoir filling.
These last two figures show the extent of internal load transfer. There is a large
transfer of stress from the soft clay core to the relatively stiff rockfill shells. In this
dam the rockfill was a very hard monzonite. The movements in it were much less
than those in the core. The chain-dotted lines indicating the overburden pressure are
based on the actual unit weights which for the core are some 75% of those for the
shells.
The vertical stress distribution across the core is of crucial interest. Safety against
hydraulic fracture will be assured if the effective stress components normal to a
potential fracture plane are compressive. The reservoir full condition is the critical
case. The results for it are given in Figure 29. The c.p.c.s.m. prediction is
reassuring in that it shows compressive effective stress across the full section whereas
the elastic analysis indicates near zero effective stress at the upstream edge of the
core. The core is, in fact, plastic so these results serve to illustrate the limitations
of elastic modelling which would appear to be misleading in this respect.
It can be seen from Figure 27 that the computed pore pressures agreed well with
the measured pore pressures until about the middle of impounding after which they
were underpredicted. The C.p.c.s.m. prediction is a little better than the elastic. A
reason for this is thought to be the ability of the c.p.c.s. model to simulate negative
dilatancy, i.e. shear induced excess pore pressure associated with a positive value of
the pore pressure parameter A. Elastic modelling cannot do this. Some of the pore
pressure rise in the latter stages of reservoir filling may be due to this factor. (One
cannot be certain because there are other factors which could cause the rise.)
331

~= El in metres

y I~------__~I 50
________ ~
100
________ ~
150
________ ~
200
________- L________- L______
250 300
~~.
metres

Figure 26. Monasavu dam - finite element mesh (after Naylor et al. 1988)
332

1981 1982 1983

0
5. -0.5
04-
c::
E -10
~ -15
.....
Q.J
(/)
-2.0
4
Construction
~4 Impounding
.. 14 Operation
..
Computed a - EI 730 Measured El 730 Elas tic
b - EI 705 6 El 705 epes .m
c - EI 680 o El. 680

..
~-----
Construction Impounding
-------~~------------~~1~4r-~--------~
Operation
..
750
E 740
c::
.....ro 730
0

> 720
Q.J
W
710
700

Q.J
CL

14 J.. 1982 1983

Computed a - El 730
b - El 705
Measured
El 730 _ _ Elastic
!J El 705 - - - cpcs.m
c - El 680 o El 680

Figure 27. Monasavu dam - measured and back-analysed time v. settlement and
piezometric elevation
333

Upstream (are Downstream

2000

Overburden (1ih)
V1
V1
eu
./
.......
.~
.....
L

V1

nJ 1000 "

.....u
L
stress
eu
>

o
100 150 200 m

Legend Linear elastic


- - - (P ( S model
6 Measured total stress
o Measured pore pressure

Figure 28. Monasavu dam - end of construction total stress and pore pressure on
horizontal section at el. 680 (after Naylor, 1990)
334

Upstream [are Downstreall

2000

Overburden (1!h)

./,
ro
~
.J<:

VI
VI
QJ

.....VI'-
1000
ro
.....~
'-
QJ
>
Effective stress

o
100 150 200 m

For Legend see Figure 28

Figure 29. Monasavu dam - after impounding total stress and pore pressure on
horizontal section at el. 680 (after Naylor, 1990)
335

References

Burland, J .B. (1965)


'Some aspects of the mechanical behaviour of partly saturated soils', Proc.
Int. Research and Engng. Conf. on Expansive Clay Soils, Australia,
Butterworth, Vol. 1, 270-278
Cathie, D.N. and Dungar, R. (1978)
"Evaluation of finite element predictions for constructional behaviour of
rockfill dam", Proc. Instn. Civil Engrs., Part 2, 65, 551-568.
Clough, R.W. and Woodward, R.J. (1967)
"Analysis of embankment stresses and deformations",
Proc. Am. Soc. Civil Engrs., 93, No. SM4, 529-549.
Coxon, R.E. (1986)
'Failure of Carsington Embankment', Report to the Secretary of State for the
Environment, H.M.S.O.
Dounias, G.T., Potts, D.M. and Vaughan, P.R. (1988)
'Finite element analysis of progressive failure : two case studies', Computer
and Geotechnics 6 (special issue on Embankment Dams), 155-175.
Duncan, J.M. and Chang, C-Y. (1970)
'Nonlinear analysis of stress and strain in soils', Proc. Am. Soc. Civil Engrs.,
96, No. SM5, 1629-1653.
Eisenstein, Z. and Law, S.T.C. (1977)
"Analysis of consolidation behaviour of Mica Dam", Proc. Am. Soc. Civil
Engrs., 103, No. GT8, 879-895.
Eisenstein, Z. and Naylor, D.J. (1985)
Static analysis of embankment dams, ICOLD Bulletin 53.
Eisenstein, Z. and Simmons, J.V. (1975)
"Three-dimensional analysis of Mica Dam", Criteria and Assumptions for
Numerical Analysis of Dams, (Univ. of Wales, Swansea), 1051-1069.
Ingold, T.S. (1979)
"The effects of compaction on retaining walls", Geotechnique 29, No.3,
265-283.
Knight, D.l., Naylor, D.J. and Davis, P.D. (1985)
'Stress-strain behaviour of the Monasavu soft core rockfill dam : prediction
performance and analysis', 15th ICOLD (Lausanne), 056, R68, 1299-1326.
Knight, D.J., Worner, N.M. and McClung, J.E. (1982)
'Materials and construction methods for a very wet clay core rockfill dam at
Monasavu Falls, Fiji', 14th ICOLD (Rio de Janeiro), Vol. 4, 055, R17,
293-303.
Kulhawy, F.H. and Duncan, 1.M. (1972)
"Stress and movements in Oroville Dam", Proc. Am. Soc. Civil Engrs., 98,
No. SM7, 653-665.
Maranha das Neves, E. and Veiga Pinto, A. (1988)
'Modelling collapse on rockfill dams', Computers and Geotechnics 6 (special
issue on Embankment Dams), 131-153.
Marsal, R.l. and Ramirez, L.A. (1967)
'Performance of El Infiernillo Dam, 1963/65', 1. Soil Mech. Fdn. Div.,
ASCE, 93, No. SM4, 265-289.
Naylor, D.l. (1985)
'A continuous plasticity version of the critical state model', Int. J. for Num.
Meth. in Engng., Vol. 21, 1187-1204.
336

Naylor, D.J. (1988) (Guest Editor)


Computers and Geotechnics 6 (special issue on Embankment Dams)
Naylor, D.J. (1990)
'Static analysis of embankment dams : a finite element perspective', Dam
Engineering, Vol. 1 (Issue 2), 79-99.
Naylor, D.J., Knight, D.J. and Ding, D. (1988)
'Coupled consolidation analysis of the construction and subsequent performance
of Monasavu Dam', Computers and Geotechnics 6 (special issue on
Embankment Dams), 95-129.
Naylor, D.J., Maranha das Neves, E., Mattar, Jm., D. and Veiga Pinto, AA
(1986)
'Prediction of construction performance of Beliche dam', Geotechnique 36,
No.3, 359-376.
Naylor, D.J. and Mattar, Jnr., D. (1988)
"Layered analysis of embankment dams", Numerical Methods in Geomechanics
(ICONMIG, Innsbruck), Balkema, Vol. 2, 1199-1206.
Naylor, D.J., Pande, G.N., Simpson, B. and Tabb, R. (1983)
Finite Elements in Geotechnical Engineering, Pineridge Press Ltd., Swansea.
Naylor, D.J., Stagg, K.G. and Zienkiewicz, O.C. (1975) (Editors)
Criteria and Assumptions for Numerical Analysis of Dams, (Univ. of Wales,
Swansea).
Naylor, D.J., Tong, S.L. and Amir Shahkarami, A. (1989)
'Numerical modelling of saturation shrinkage', Numerical Models in
Geomechanics (NUMOGIII), Elsevier, 636-648.
Penman, AD.M. (1977)
"The failure of Teton Dam", Ground Engineering, 10, No.6, 26-33.
Penman, AD.M. (1982)
General Report Question 55: "Materials and construction methods for
embankment dams and cofferdams", 14th ICOLD (Rio de Janeiro), Vol. 4,
1105-1228.
Penman, AD.M., Burland, J.B. and Charles, J.A (1971)
"Observed and predicted deformations in a large embankment dam during
construction", Proc. Instn. Civil Engrs., 49, 1-21.
Penman, AD.M. and Charles, J.A (1973)
"Constructional deformation in rockfill dam", Proc. Am. Soc. Civil Engrs.,
99, No. SM2, 139-163.
Potts, D.M., Dounias, G.T. and Vaughan, P.R. (1990)
'Finite element analysis of progessive failure of Carsington embankment',
Geotechnique, 40, No.1, 79-101.
Skempton, AW. and Coates, D.J. (1985)
'Carsington Dam failure', Failures in Earthworks, Inst. Civil Engrs., London,
203-220.
Smith, I.M. and Hobbs, R. (1974)
"Finite element analysis of centrifuged and built-up slopes", Geotechnique, 24,
No.4, 531-559.
337

Appendix Derivation of Equivalent Layer Stiffness

Consider the placing of depth H Q of fill where H Q is the depth compnsIng a single
layer in the analyses but in reality is constructed in a large number, n, of sub-layers.
Let b represent the displacement of a typical point on the layer base. Referring to
Figure AI, let w = ')'HQ be the gravity loading intensity due to the weight of the
new layer, q be the pressure transmitted from the new layer to the underlying
material, and p = w-q. If the new layer has no bending resistance then q = wand
p = O. In general, p will not be zero, but its integral over the full area of the
layer base must be zero to satisfy equilibrium. p will esentially depend on the
flexural rigidity (EI) of the new layer. (The shear rigidity of the layer may also
make some contribution but, as in simple bending theory, this contribution will be
neglected. )
Restricting attention to linear elastic material, the object is to find an equivalent
flexural rigidity EI (with corresponding equivalent Young's modulus) for the analytical
new layer which will give the same displacement, 5, and pressure, q, on the layer
base (and therefore the same p) as would be produced in the actual case in which
the fill is built up in n sub-layers.
To do this consider an arbitary length along the layer, e.g. over the central part of
the embankment, as shown in Figure A2. Let the forces equal to the pressures
integrated over this length be denoted by capital letters, i.e. P. Q and W. and let p
be a measure of the average settlement. Let So be the vertical stiffness of the
underlying fill, and define a stiffness S by

S = Pip (AI)

S can be interpreted as the vertical stiffness of the chosen arbitrary length of layer
due to its flexural rigidity. It is assumed that S is proportional to the flexural
rigidity, EI, of the new layer. Denoting the value of S for the completed layer with
its stiffness unreduced by S' the stiffness when m out of n equal sub-layers have
been placed is then

'l~
I I~
I I
I
I
__ : I
"\J
vertical pressure distribution underlYing fill

Figure AI. Loading on new layer Figure A2. Forces on arbitrary length
33R

CA2)

This assumes equal layers with lin times the total load W applied to each of them.
During the placing of layer m let P increase by .1P m and p by .1Pm' Therefore

Summing for the n layers

n
P = L Sm.1Pm CA3)
m=1
At this stage it is necessary to make an assumption about the rate of accumulation
of settlement as the new layer is built up. If the effect of the layer flexural rigidity
is relatively small, i.e. S < < So then P will be small so that the displacement is
almost entirely due to the weight of the new layer transferred directly on to the
underlying fill. Therefore

i.e . .1Pm = constant = pin


Equation (A3) then becomes

P = !!.
n

Introducing (A2)

pSI
n
P
f14 2: m3
m=1
On summing the series(i)

P = pSI (1 + ~)2 (M)


4 n

Elimination of Pip by Equation (A4) now gives the equivalent stiffness, S, as

(AS)

Taking n as large the stiffness reduction factor is then

f 5'/5 = 4 (A6)

n
4n
4
(i) 2: m3 (1+n)2
m=1
339

If the contribution of the flexural rigidity of the new layer is so large that it is
unreasonable to assume S < < So (or P < < W) then the rate of settlement will
vary as the layer is built up. When the first few sub-layers are placed the effect of
the flexural rigidity will be very small since EI is proportional to the cube of the
layer thickness (c.f. Equation A2). The effect will increase rapidly as the layer
approaches completion. In regions of the fill where sagging bending moment is
induced in the new layer (as in the centre of the fill of Figure AI) the increasing
flexural stiffness as the fill is built up will tend to inhibit fill settlement. Q will
become less than W, P will be positive. The initial displacement increments will then
be greater than the subsequent as illustrated by the curves S l ' S 2 of Figure A3.
Conversely if the bending moment is hogging P will be negative and the effect of the
flexural rigidity will be to increase the settlement. Curves such as HI' H 2 in Figure
A3 will then be followed.
To assess the effect of this the reduction factor is calculated making the assumption
that the settlement of the new layer case varies quadratically with the increasing fill
height in the new layer. This is achieved by expressing the settlement due to
sub-layer m as

A+ B m (A7)
n

A and B are constants for a particular curve. A family of curves of this type are
defined by setting

ap
A
n

and

2(I-a)p
B
l+n

"a" defines the particular curve, and has the values 0, L I, 3/2, 2 for the respective
curves H 2' HI' L, S l ' S 2' (nLn is the linear relation for which the reduction factor
of 4 has been derived.) it is readily shown that the summation of .1Pm for n
sub-layers is p for all "a" values.
Introducing Equation (A7) into (A3) now leads to

n
. 2(l-a)m
p
2 (~)3 5'
n n [a + l+n
m~1

n n
p5'a 2p5' (I-a)
fi4 2 m3 +
n 4 (I+n) 2 m4
m~1 m~1

Summing the series(i) gives

n
(i ) n
2 m4 30
(l+n)(1+2n)(3n 2 +3n-l)
m~1
340

Pal (I-a) 1 3 1
pS' =4 (1 + n)2 + -1-5--- (2 + n)(3 + n - ~) (A8)

Eliminating P/ p by (AI) and setting n large one obtains

5'
f = S (A9)

Values of the reduction factor, f, corresponding to the five curves are marked on
Figure A3. Note that f = 4 is reproduced with a = 1. There is an upper limit on
"a" of 8/3 when the factor becomes infinite.
In conclusion it has been shown that reduction of the final (in situ) stiffness of
each analytical new layer by a factor centred on the value four will give a bending
stiffness for a single step application of gravity equivalent to that of a continuously
place fill. Linear elasticity has been assumed.

1D

hogging

fraction of
new layer
placed (minI
0.5

0.5 10
fraction of settlement, p

Figure A3. Effect of new layer bending made on development of displacement and
optimum load factor
CHAPT ER 13
CONCRETE FACE ROCKFILL DAMS
NELSON L. de S. PINTO

I. INTRODUCTION

The concept of a rocki! ll d am with an impervious fac e dates from the


California gol d ru sh of the 18605 and 18708 [11. The int uitive use by
th e mln~r s of easily ava il able ~aterlals (rock and lumber) a nd their
know-how in roc k blasting works le d ofeen l y to the const ru ction of
timber face rockf!ll dams , to s t o re water for the sluicing operations .
Rei nfo r ced conc rete slabs r ep l aced timber as the impervious element
for higher and pe r manent dams in irrigation and power proj ec t s . and the
concre t e face rockiill dam (eFRO) . soon became a recognized and
respected dam t ype alte r na t ive . A s trong and reliab l e s tru ct ura l body.
f r ee o f upl ift or internal pressure pr oblems , a n ideal l y loca ted
impervious plane at the upst r eam f~ ce , t ogether with s i mple and
straightforward construction procedure s , were the ma in a ttributes o f
thi s new type of dam .

Dumped CFRDs became popular i n the first half of th is ce ntury .


Actually very few impervious core r oc kf il l dams existed prior to 1940.
As the hei ght of the dams increased an d nea r ed 100 . 0 m, l eakage due t o
high fill deformation s and ope ning o f the jo i n t s in the concrete face
became a frequent problem . Although neve r a safety question , l eakages
were a nuisance . enough t o divert the selection of the dam t ype towards
the ea r th co re r ockfi ll. which beca~e the dominant r ockfi ll dam t ype for
the iollowing 30 year s [2].
Compac t ion of r oc kf i11 dams s ta rted in the 1960s , r esulting in a
much less defo rmable fil l, mor e comp"tible with the needs fo r an
impe rvious conc rete membrane. Join t design was perf ect i oned and
l eakages con t rolled to very reasonable levels . Gr ad ually th e CFRD
resumed it s place among r oc kfill dams . Found ation r e quirements being
e sse ntially the same as f or the cen tral core dam, CFRD's ot her
at t r ibut es suc h as simpler cons truction logistics, less cos t, more
compa c t layout , easier rive r handling solutions , short e r const r uct i on
t ime, have been we ighing i n it s f a vor . Indica tions a re thst CFRD is
be ing c hosen more and more frequently i n the last decade [3J .

'"
E. Maran~ daJ N~l'tJ (~d.J. Ad\'(J~tJ in Rodfrll Slr"CI ..r~J. 341-373.
CI 1991 KlllWtr Academic Publishers.
342

Sec t I on

~~
Longitudinal Section

Early CFRD - 23m high ESCONDIDO DAM


CALIFORNIA - Circa 1890

At 60' @ C~c;ed a~,s


v,,,,,,,o,,,~~,
@ HQrizontal Jo'nl~ _~
CD \*i20'r Moslle
Cutoff Concrete foce ~ 4 15' Premoided

<:#
(Trench) ~ A!>pholt
I 3A-m reQwoocl ELEVATION OF FACE /,
fdler and ~ CompreSSible
..)9 UCopper Jo,~' I,'ler
... Z wal.mop 13+(Nalurol)
@ ~I r-::. -.-i
,,' 1:5
I

Hf;;~'OI Crone-placed @ Vertlcoj JOon'


lor(je rOCk

...
~d:: 4

SECTION OF DAM

Typical traditional design of dumped CFRO - 1940s

~.~/
- ... - -.--.,-~
li""'~"" (osdeterm"1edb~c~trcclor)
v",,,,, 1""--- ~mnwl'GHTAX"
@Hor,zontolconstrucllonja,nl - I I : !
: I ! :
[}owe,: I ~ f~:~~~ CD ~ 20r- ,I 0
Groutlone~ G:l -
As."" Concretl foce
rom",,.,
ELEVATION OF FACE Asphalt_pOInted
"IdIOio'

~(7,~< ~I@ ~
"" Woterslop,
HOrlZol'1lal ~"".,".tlberOrVInYI
cOl"l51ructlon Compocted or dumped \::::.J
jOint rock fill -<!I Verl,eol
""..../)<1 JOint
"I 40"
I--Dowel~
..... Grou! I,ne SECTION OF OA M

Typical recent deSign af compacted CFRO

Figure 1. Evolution of the CFRD concept [1], [4].


343

The excellent performance of recent very high dams, like Foz do


Areia (160 m) and Salvajina (148 m) testify to the degree of acceptance
and reliability of the CFRD. Several dams in the 180-220 m range are
presently at different design stages, and consideration is already
given to the 300.0 m mark, up to now an exclusive realm of the
impervious core dam.

2. CURRENT DESIGN PRACTICE

2.1. Evolution

Modern CFRD design practi.ce is the result of experience. Basic


dimensions, rockfill zoning, constructional details are all based on
successful precedents. Most commonly, no stability or structural
analyses are carried out. Design innovations are to be considered with
great care and only after a through knowledge of the history of the
design practice and of the reasoning behind the many practical details.
A most comprehensive picture of the history and development of the
CFRD dam and of its design is found on references [4], [5], and [6] in
which the contribution and influence of J. Barry Cooke cannot be
overemphasized. A glimpse over the main aspects of that evolution and
of the present status of design is given below.
Early CFRDs were all dumped concrete face dams. By the 1940s,
their design followed some well established principles, and their main
features could be resumed as follows.
Fill - Dumped rockfill in 5.0 to 60.0 m lifts. The upstream face made
smoother to receive the concrete face by a layer of derrick
placed rock blocks.
Cutoff - The face slabs were made to rest on a concrete filled cutoff
trench excavated in the foundation rock.
Concrete face - Relatively thick face slabs (e = 0.3 + 0.0065 H meters),
divided by vertical and horizontal soft joints. Steel
reinforcing of 0.5% of the concrete design section in both
directions at the center of the slab.
Joints - Horizontal joints 1.0 to 2.0 cm wide and vertical joints 2.5
to 5.0 cm wide, provided with copper waterstop and separated by
compressible fillers - pre-molded asphalt and soft wood planks,
Face slopes - Natural dumped rock slopes, most generally 1.3H:1V.
Axis - In an attempt to reduce the tendency for opening of the vertical
joints, a slight curvature concave upstream was adopted.
Slab settlement under water load was found to be essentially normal
to the face with the pattern illustrated in Figure 2, maximum deformation
taking place at about 0.4 of the dam height. As the concrete slabs
follow the rockfill movement, they are dragged towards the center of the
dam, the joints closing at the center and opening near the abutments.
For dams in the 100.0 m range in height, joint movements were quite
large and excessive leakage occurred, due to joint openings, spalling
at joints, or cracks in the concrete face.
344

~~~~+~-'T-~~- 5800

o ~O 100FT ,V / ~~~~~~-5650

'=='==J \" ///\


SCALE --~
-~~~=-~~=-'----~~~~~~~~~-5600

A B

Figure 2. Slab deformation - Lower Bear River Dam No.1 [7].


A - Settlement contours (1) Vertical joint
B - Main section of the darn (2) Horizontal joint

For compacted embankments the overall picture of deformations and


face movements is the same but with values of much smaller magnitude.
The modulus of deformation of dumped rockfill is five to ten times
smaller than those attained in compacted rockfill where values of 50 to
150 MPa are commonly reached depending of rock gradation, layer
thickness and compaction effort.
A further step towards a less deformable dam fill is the
replacement of the derrick placed rock blocks layer at the upstream
face by a transition zone of compacted finer and well graded rockfill
material. Compacted in thinner horizontal layers and along the face
proper by a smooth drum vibratory roller, the transition zone results
in a low deformability, semi-pervious layer with an ideally continuous
and regular surface to support the concrete slabs.
The concrete face could consequently be made thinner (e = 0.3 +
+ 0.003 H meters). Steel reinforcing was initially maintained at 0.5%
in both ways, with a tendency to be reduced as no large gaps
between rock blocks exist to be bridged as in the ancient design.
Horizontal joints were suppressed and slipform construction adopted as a
pratical and economical concrete placement method. Vertical cold joints
without soft fillers help to minimize slab movements and joint openings.
The concrete cutoff wall is replaced by a toe slab, or plinth,
dowelled to the rock, avoiding the deep trench excavation. The face
slabs float on the rockfill with no restraints from the plinth.
The supposedly favorable arch effects of the curved axis of the dam
was recognized innefective, and practically all modern CFRD dams adopt a
straight axis.

2.2. Embankment

2.2.1. General comments - The main scope of CFRD's design is to make


the relatively deformable rockfill compatible with the rigid concrete
345

slabs of the face, for the joints to be able to absorb the rockfill
settlement without undue leakage.
The deformation of the dam depends on its height and on the
modulus of deformation of the rockfill. The load is the result of the
dam weight and of the water pressure against the face. The resulting
settlement is essentially due to the rupture of the rock blocks at
their points of contact and rearrangement of the rock particles. Its
intensity is relatively high immediately following the application of
the load and reduces gradually with time. Total settlements can be
assumed proportional to the square of the dam height and inversely
proportional to the modulus of deformation. The modulus is higher, the
sounder and better graded the rock and smaller the resulting void ratio
of the compacted fill.
Deformations due to the weight of the rockfill occur mostly during
construction and have pratically no influence on the behavior of the
face. Ideally the concrete slabs are placed after completion of the
embankment and will be essentially influenced by the settlement due to
the water load. In high dams it may be practical to raise the concrete
face to an intermediate level prior to the completion of the fill. It
has been proven in several cases, that deformations of the embankment
due to the added weight of the upper reaches produce no harm to the
lower concrete face, basically because no hydraulic pressure acts on
the slabs which simply follow the movements of the fill.
The settlement due to water load depends mainly on the deformation
properties of the upstream portion of the dam. The downstream shell
complements the dam section but its influence on the movement of the
face is irrelevant. The zoning of the dam and respective compaction
requirements are consequently more rigid for the upstream third or half
of the dam section, (finer layers, better compaction). For the
downstream shell thicker layers and/or dumped rockfill are accepted.
The rockfill zoning of Foz do Areia dam is compared to the zone
designations for CFRD suggested by Barry Cooke and Sherard [6] in
Figure 3 and used as a reference to further comments on the role
and basic aspects of the different zones of the embankment.

2.2.2. Zone 1 - Impervious blanket - An impervious blanket over the


plinth and the lower part of the concrete face has been used in several
high dams, as an extra protection against possible leakage at this most
critical zone. As it will be commented under "joints", this secondary
protection has been receiving closer attention recently and may be made
to represent a basic line of defense against leakage in very high dams
where waterstops may not resist the large joint movements.

2.2.3. Zone 2 - Processed small rock transition - The transition zone


in Foz do Areia has a maximum width of 10 m at the foundation, tapering
to 4 m at dam crest. A constant width of 4.0 to 5.0 m is now generally
accepted. Widening to 10.0 to 20.0 m near the foundation recognizes
the more critical zone and helps to minimize settlement below the
perimetric joint. The grading curve for Foz do Areia dam is shown in
Figure 4. The crushed run material limited to 12.5 em diameter and
compacted in 0.4 m layers was somewhat short in fines. Current practice
346

calls for a finer (minus 7.5 cm) better graded material with sufficient
amount of sandsized particles to act as a filter to control eventual
leakage paths through cracks or joints, mostly near the abutments, and
to effectively act as a semi pervious upstream zone in case of
occasional floods during construction, prior to placing of the concrete
face,as illustrated by the material II-BB specified for the region of
the perimetric joint in Segredo dam under construction.

ANGLE DEPENDS ON HEIGHT

TABLE OF MATERIALS
MATERIAL ZONE METHOD OF PlACEMENT
IA IMPERVIOUS SOIL 38 qUAR,,?'( RUN AOCKFlll,
DUMPED ABOUT I m LAYERS
18 RAN DOM
3C QUARRY RUN ROCK FILL,
2 PROCESSED SMALL ROCK ABOUT 15 TO 20m LAYERS
1--_---+'-"-1"::::""::::"::::"::::"::::"::::'::::"::::"::::"::::"::::".l------
~T~~~~::~A I D COMPACTED1NOaOmLAYERS SELECTED SMALL ROCK
PLACED IN SAME LAYER
(s.:;~~~~~~~S:;:m'nl I E PL"'CEDF!OCK(~'Ixe) THICKNESS AS ZONE 2
WELL GFlAOED_MAX_SIZE 6" LIlYEAS _4 PASSES OF VIB ROlLUI
II B COMPACTED IN LMERS 040 m FACE -5 PASSESVIB ROLL(R(WWVe)

[AIlHIFllL ~A~IMUM SIZE3I4"COMPACTUI PNEU'-IATIC ROLLER OR

'"
1110 IN 030m L.o.Y(RS CONSTRUCTION EQUIPMENT

A 8

Figure 3 - CFRD rockfill zoning


A Foz do Areia darn (2) El. of first stage
B Zone designations (3) Parapet
according to [ 6] (4) Crest
(1) Soil protection (5) Dike

The transition material is generally compacted in 0.4 to 0.5 rr, thick


layers with 4 passes of a 10 t or slightly lighter vibratory roller and
along the face by 4 to 6 passes of the vibratory roller, vibrating when
driving up slope, to avoid ravelling of the face. Backhoe-mounted
plate vibrators have been successfully used for face compaction and
are particularly useful to achieve good compaction in restricted areas
near the plinth.
The face of the transition material should be protected with a
sand-asphalt emulsion or gunite layer to reduce the risk of erosion
due to heavy rain, particularly in the zones of concentrated runoff
along the abutments.
347

200 40 10 4 6"
100 --' o
II II /1/ /
80
/ / / II I 20
II / I II
r / / _(2) /
l v 40 ~
1l2J-1'- e-.. II / 1/
'" 60
~
t>Z
c a
II o
u
~
o 40 / N / ~ I------ V cf!.
7 II
60

V I/. ij-- / "XI)


v / vv ...... ~ V
~v
/
20
v L,.-V
80
/V l--------
o
0.01
/

0.02
v-
0.06 0.1
---
0.2
I----- ~
0.6
-f--
2
Diameter of particle in mm
v
6 10 20 60 100
100
200

Figure 4. Transition material grading curves


(1) Granulometric envelope of lIB transition, Foz do Areia dam
(2) Granulometric envelope of IIBB transition, Segredo dam
(3) Silty sand material

2.2.4. Zone 3 - Main rockfill embankment - The main rockfill embankment


includes zones 3A, 3B and 3C. Zone 3B of lowest compressibility
upstream is generally compacted in 0.8 to 1.0 m thick layers, water
being added at 15% to 25% of the rockfill volume, maximum rock size
limited to the thickness of the layer. A transition layer, 3A, has
been adopted lately following the practice of finer zone 2 transition,
with limited rock block size, compacted at the same time and in the
same thickness as zone 2, to provide a gradual transition to the main
rockfill body. It was not provided in Foz do Areia dam. Zone 3C
completes the downstream shell where a more deformable embankment is
acceptable. Usually compacted in 1.5 to 2.0 m thick layers, zone 3C
accepts also dumped rockfill. A final layer of large blocks dozed to
the face is good practice for a neat and safe downstream face.
The rockfill grading is generally accepted as produced in the quarry.
A well graded material is bound to result in a denser and less
deformable rockfill, but less than ideal gradings have proven
satisfactory also. The range of gradation recorded in Foz do Areia dam,
where a sound basaltic rock produced a well drained rockfillparticularly
short of fines (uniformity coefficient - 0.6), is illustrated in Figure
5. Relatively low deformation moduli of about 50 MPa for the 3B zone
and 25-30 MPa at the downstream shell resulted with no adverse consequences.
348

The free draining characteristics of the rockfill are further enhanced


by the zoning. The thickening of the compacted layers froID zone 2
through zones 3-A, B, C, results in a favorable permeability gradient.
Acceptable limits as for the amount of fines in hard rock rockfill
are indicated in Figure 5. When dirty rock is used with excess fines
which could result in a less pervious fill, interior drainage zones can
be provided to assure a safe free draining embankment. As for the
strength, rockfill with excess of fines is often evaluated by the
trafficability when throughly wetted. Deformation under loaded trucks
will detect a soil like behavior if fines are off limits for an
acceptable rockfill. Rockfill with up to 60 - 80% minus 25 mm, 40% minus
5 IT@, and 5% minus 0.1 mm is being accepted in selected regions of the
3C zone in Xingo dam under this type of control.

200 40 10 4 6" 12"


o
II
100

/ /
80 --,I / / 20

G: i / /
[7
60 /-- ~ ;---.... / 40 ~
l.
J I------ J
~

'" I2
\,~
/ o
o

cf!. 40 ---------- I-~ 17 [7 7 I


U

cf2
/ /
60
/
V
-
l--' /
./
,-- .-I- 1-'--
20 80

.- ~ 1-1--- \---V V
V
-
f--
!- ,-f---
o
0.06 0.1 0.2 0.6 2 6 10 ----
20 60 100 200 600
100
1000
Diameter of particle in mm

Figure 5. Foz do Areia Dam - Rockfill grading as recorded in the


field (117 samples)
(1) Granulometric envelope of IB rockfill
(2) Acceptable limits for fines in hard rockfill

2.2.5. Fill cross section - For the dumped CFRDs, face slopes at the
natural angle of repose of rockfill were accepted with no questioning
of their stability. There is no reason to change that practice in
compacted dams. An average slope of 1.3H:1V is generally used for most
dams, while a 1.4H:IV slope is justified for dam heights above 120 m.
A 1.3H:IV slope corresponds to a safety factor greater than 1.5 for a
o = 50 0 rockfill based on the infinite slope method. The flatter slope
for very high dams recognizes the lowering of the friction angle as
normal pressures increase, as pointed out by Leps [8]. Somewhat flatter
slopes (1.5:1 to 1.6:1) result for gravel fills. Berms or roads in the
349

downstream face may produce acceptable steeper intermediate slopes of


about 1.2:1 or 1.25:1 when finished by selected bigger blocks as in Foz
do Areia and other dams.
Crest widths of 8.0 to 12.0 m have been used, the main limitation
being the space needed for the concrete works for the face.

2.3. Plinth

The plinth or toe slab is at a time a grout cap for the grout curtain
of the dam and a watertight articulated connection of the face to the
foundation.
Classical design is a thin reinforced concrete slab, dowelled to
the rock as illustrated in Figure 6. The plinth is adapted to the
topography of the site in straight continuous stretches without joints
but for construction joints with reinforcement passing through, spaced
at the convenience of the contractor. For good non erodible foundations,
the cross section length is generally made to be from 1/10 to at least
1/20 of the hydraulic load. Thickness may be the same or less than for
the face slab and reinforcement is minimum temperature steel. On steep
abutments, it can be designed as nearly vertical walls anchored against
the rock to minimize excavation. At Foz do Areia and Segredo dams, the
plinth foundation was designed as a road, sometimes deeply excavated in
the more steeper abutments, with horizontal generatrix normal to the
perimetric joint. The resulting larger excavation is compensated by
advantages of a better control over rockfill thickness under the
perimetric joint, a more straighforward concreting scheme, and greater
facility for moving the crawler mounted drilling equipment and for
drilling the grout holes. Somewhat high plinth sections may result from
abrupt changes in the slope of the abutment. In those cases the
structure has to be checked for stability taking in account the
unfavorable conditions of high horizontal and uplift pressures.

Figure 6. Segredo Dam. Plinth, typical cross section


(1) Hard groutable rock (5) Slab
(2) Dowels (1,2 m both directions) (6) Mortar pad
(3) Grout curtain holes (7) Base concrete
(4) Consolidation grouting holes
350

If danger of erosion exists in softer seams in the foundation, an


extension of the seepage path to one fourth of the water head or more
can be provided by shotcrete lining of the foundation downstream of the
plinth, dully protected with filter layers to avoid the migration of
fines into the rockfill.
A most impressive and instructive example of the possibilities of
the plinth concept extended to weathered rock conditions and to residual
soil type of foundation is the Salvajina project [9], illustrated in
Figure 7.

',,*,;;;;j~::::-t- VERTICAL JOINT

TRANSVERSE JOINT

PERIMETR1C JOINT

A - A
FOUNDED ON COMPETENT ROCK
(Foundation type I)

_7_-~lo~m~---t'1 PERI METRIC


JOINT
(CONNECTING SLAB ZONE 1

~~~>;='i==n~M--=:\ I _(~

"~:N~ii
ZONE 2/ I

FOUNDED ON LESS COMPETENT ROCK


(Foundation types II and III)

Figure 7. Plinth at Salvajina Dam [9].

Grouting is carried out along the plinth, outside the rockfill body,
and in no way interferes with the construction of the dam. It can be
performed meticulously without affecting the overall schedule which is
a very important feature. A high hydraulic gradient resulting from the
short width of the plinth and the lack of dam weight directly above the
grout curtain recommend a careful grouting program.
The toe slab functions as a very effective grout cap dowelled tothe
rock. Upstage packer method is usually employed. Two rows for shallow
treatment upstream and downstream of the curtain line and a deep central
curtain grout make up a common and adequate scheme. The consolidation
shallower treatment should be regarded as an important item, as it
checks the fissures in the more critical high hydraulic gradient
foundation area.
351

Sound groutable rock is the ideal plinth foundation. More complex


foundations lllay require special treatment and design solutions. In
Khao-Laem dam, in Thail nd, the toe slab was combined with a diaphragm
wall to successfuly control seapage in a karstic foundation [12].

2.4. Concrete face

2.4.1. Slab thickness - Concrete face thickness defined by e = 0.3 +


+0.003 H m, is accepted without questioning. A clear tendency exists
for thinner slabs, as in Tasmania where constant 0.25 m thickness is
preferred for moderate height (40 to 75 m) dams. For higher more
important dams e = 0.3 + 0.002 H seems to be a reasonable compromise.

2.4.2. Concrete - Strength is not as important as durability and


impermeability. For Foz do Areia,pozzolanic cement 310 kg/m 3 was used
for the air entrained concrete of the slab. Design strength: 20.6 MPa
at 28 days, water - cement ratio: 0.53, slump: 8 cm.

2.4.3. Reinforcing - Reinforcing of 0.4% of the design concrete section


in each direction in the center of the slab is classical. Reinforcing
is basically needed to prevent temperature cracks, since essentially
all the face is placed under compression by the water load, but for a
limited zone near the abutments where low tension values have been
observed. The face rests on a quite uniform compacted rockfill surface
and normally no moments should be transfered to the slabs. However,
near to the foundations, particularly in the case of steep abutments,
flexional and/or torsional efforts can result due to differential
settlements of the slab, and should be overtaken by the reinforcing.
For most of the central part of the face, lighter reinforcing,
0.3 - 0.35%, is being accepted.

2.4.4. Joints - Vertical joints normal to the axis of the dam are cold
joints, spaced at 12.0 to 18.0 m, in general. The spacing of the
joints is related to the dimensions of the formworks for the generally
slipformed face construction. Wider slabs result is less joints but in
heavier and more cumbersome formworks. A joint spacing of 16.0 m as
adopted in Foz do Areia was a good compromise and proved quite practical.
Vertical joints are made to end normal to the perimetric joint,
particularly in steep abutments, to avoid sharp angles at the end of
the slabs.
Horizontal joints have been totally eliminated. Construction joints
when needed are treated as such, with reinforcement passing through
and no waterstops.

2.4.5. Joint details - Details have been developed to control leakage


through the joints. Practically, the sole movement to be considered
along the vertical joints is that of opening or closing of the joint,
as rockfill deformation varies gradually and no offsets occur. The
vertical joints tend to close in the center and some openings are
observed near the abutments.
Joint design takes in consideration the two cases as shown in Figure 8.
352

Figure 8. Vertical joint details


A - Central vertical joints (3) Painted with beturnen
B - Vertical joints near the abutments (4) Mastic
(1) Concrete pad (5) Hypalon rubber cover
(2) Copper waterstop (6) PVC strip

A copper waterstop over a concrete pad in the bottom of the slab


is a proven solution for the central vertical joints under compression.
Vertical joints with passing reinforcement is an acceptable
alternative for the central zone of the face, at least for regular
shaped valleys, with or without the bottom copper waterstops.
For the lateral joints with a tendency to open, a double barrier ib
provided. A central waterstop has been used in addition to the copper
waterstop. Due to the difficulties of concreting under the central
waterstop, recent preference is towards a mastic filler over the joint,
covered by a hypalon rubber, fabric reinforced rubber, or similar
impervious membrane.
The main joint movements occur in the perimetric joint along the
plinth due to the discontinuity between the toe slab founded on rock
and the slabs supported on the deformable rockfill. In general three
components are observed: settlement normal to the face, opening
normal to the joint and a tangential component parallel to the joint,
as illustrated in Figure 9.
Maximum reported movements vary with the height of the embankment
and the construction deformation modulus of the rockfill, as seen below:

Movements <mm) Settlement Eening Tangential OBS.

F.do Areia H=160 m tangential value


E=37 to 55 MPa 55 23 25 estimated

Cethana H=110 M
E=1l2 to 185 MPa 11 7

A.Anchicaya H=140 m Around weak zone


E=98 to 167 MPa 106 125 15 at right abutment

Shiroro H=125 ill


E=76 MPa 50 30 21
353

A 8

Figure 9. Perimetric joint movement


A - Section (3) Tangential movement
B - Plan (4) Plinth
(1) Opening (5) Slab
(2) Settlement

The potential for leakage is correspondingly high and particular


care should be taken in design.
As the first compacted concrete face rockfill darn above 100 m in
height, Cethana Darn (110 m - 1970) has defined a joint design
essentially based on a double waterstop system, as indicated in Figure
10-A. The design was adequate for the slight deformation experienced
by the dam and leakage was limited to less than 50 lis, reduced to
10 lis after 5 yrs.
Alto Anchicaya Dam (140 m - 1974) followed with a simplified one
rubber waterstop concept, Figure 10-B. Localized rockfill deformations
higher than expected and problems of weak concrete around the waterstop
in some points have resulted in high leakages, up to 1,800 lis.

A B

Figure 10. Perimetric joint details


A - Cethana Darn (3) Rubber waterstop
B - Alto Anchicaya Dam (4) Concrete pad
(1) Plinth (5) Slab
(2) Copper waters top
354

Remedial works after emptying the reservoir included rebuilding


the concrete around the waterstop and filling above the joint with a
mastic. A rubber hose was also used in zones where large openings had
been detected.
The design of Foz do Areia dam (160 m, 1989) profited from both
experiences. The joint detail incorporated the double waterstop concept
of Cethana and provided, at the construction stage, a mastic reservoir
as a self healing system in case of damage to the waters tops due to
greater joint movements, Figure 11.

Figure 11. Perimetric joint, Foz do Areia dam


(1) Copper waterstop (6) Face slab
(2) Mastic (7) PVC waterstop
(3) PVC cover (8) Sand asphalt pad
(4) Neoprene tube (9) Timber filler
(5) Plinth (10) Special grading zone IIBB

The two waters tops embedded in the concrete were planned to be of


different materials and/or geometry to minimize the chances of an
eventual direct open orifice in case of failure of one waterstop; the
mastic provided the self healing effect in case of leakage; the sand-
asphalt pad was to act as a backup for the lower copper waterstop and
to prevent the flow of the mastic; the neoprene cylinder should help to
close the gap due to a greater joint opening.
The excellent behavior of Foz do Areia dam, as far as leakage is
concerned, has ratified the perimetric joint detail. Many other projects,
such as Salvajina, Cirata, Terror Lake, have similarly conceived joint
systems.
It is well understood that a convenient behavior of the mastic in
preventing leakage depends on the overlaying neoprene or rubber sheet
to distribute the water pressure and to cause the flow of the mastic
into the eventual opening.
In Foz do Areia as in several high dams the lower part of the
plinth has been covered by an additional earth fill blanket, which may
3SS

well constitute the main water barrier in this most critical region.
Certainly, a soil type material, preferably a silty-sand material as
illustrated by sand (3) in Figure 4, will have an effective sealing
effect on possible cracks and openings, if well backed up by a proper
filter underneath the slab. Such is the idea behind the proposed
perimetric joint concept [11], illustrated in Figure 12.

Figure 12. Proposed perimetric joint concept [11].


(1) Fine sand (6) Sand-asphalt or mortar pad
(2) Filter - IIBB material (7) Copper waters top
(3) IIBB with 5% cement (8) Silty fill
(4) Plinth (9) Normal transition material
(5) Slab

Present design tendencies can be resumed as follows:


- bottom copper waterstop supported by a concrete pad, which replaces
the sand-asphalt pad used in Foz do Areia;
- elimination of the central waterstop, due to difficult concrete
placing conditions and weakening of the concrete slab;
- mastic filler covered with a fabric reinforced rubber membrane
thightly tied to the concrete;
- a strip of silty sand over the joint, as an additional seal against
eventual leakage.

2.4.6. Parapet wall - A vertical parapet is a natural finishing detail


for the concrete face, both for personnel safety reasons and as a
freeboard against wave splashing. It has proven economical, due to
the resulting reduction in rockfill volume, to set the parapet wall
about 4.0 to 8.0 m in height as for Foz do Areia, Salvajina and other
recent dams, Figure 13.
356

-6.00 ,
B.OO
EL.749.DO:Min.) r EL.1IS2.00 t t
I F==t===i1
EL.IISI.SO

CREST DETAIL

A B

Figure 13. Parapet wall


A - Foz do Areia Dam B - Salvajina Dam

3. CONSTRUCTION FEATURES

3. 1. Embankment

The rock from the quarry or from required excavation is brought to the
embankment by end dump off-road trucks, spreaded by dozers, and compacted
by smooth drum vibratory rollers. For sound rock, the material is
dumped on the layer being compacted and spreaded, to advance over the
previous layer, as illustrated in Figure 14. A degree of segregation
is typical of this process and is accepted as a normal feature of the
rockfill. The lower more permeable layer is formed by larger blocks,
and the denser better graded top zone smoothed by the vibratory roller
results in a favorable surface for the traffic of the construction
equipment.
For weak rocks, the dumping can be made directly over the
preceding layer, reSUlting in less segregation and better overall
density.
Water at a ratio of 0.15 to 0.25 of the rock volume should reduce
the resistance of the rock to breakage and result in a higher degree of
compaction. It is generally required for the 3B zone. Water ulay not
be specified for the downstream, 3C shell, as no major concern exists
over the resulting deformability of the embankment. High pressure
monitors are not needed as the aim is simply wetting the material and
not moving the rock fragments.
Compaction by four passes of a 10 ton smooth drum vibratory roller
is standard. Particular conditions may recommend a larger roller or
additional number of passes.
Construction joints within the embankment are a frequent feature
of CFRDs and one of their great assets as far as construction logistics
is concerned. Both longitudinal or transversal joints at a natural
1.3H:IV slope are acceptable anywhere in the body of the dam, except for
357

the first 20.0 to 30.0 m downstream from the face, where specifications
should require construction in horizontal layers well placed and
compacted. Construction roads and ramps within the rockfill, with
slopes up to 18%, as well as a permanent road along the downstream face
of the dam are frequently used for hauling of the rock, Figures 15 and
16.

Figure 14 - Rockfill placement at Segredo Dam.

The finer transition zone 2A is compacted in a natural damp


condition. Excess water 'is avoided as it may tamper the compaction
effect. For compaction along the face, a 6 to 10 ton roller, vibrating
up the face, has been mostly used. Four to six passes are the normal
procedure. An asphalt or shotcrete layer over the face improves the
stability of the fine material against ravelling. When a backhoe-mounted
plate vibrator is used for face compaction, its effect is controlled by
the settlement, usually specified around 7.0 to 10.0 cm. The plate
vibrator is particularly suited for compaction of the critical zone
close to the perimetric joint where good compaction by the vibratory
roller is oftenly impossible.
Protection of the waterstops partially embeded in the toe slab is
needed during compaction. Heavy timber planks may be placed around the
waterstops, arranged in such a way as to preform the space underneath
for the casting of the mortar pad also. A small amount of cement, say
5% by weight, added to the transition material near the joint will
result in a cohesive mixture and avoid loosening of the fill in the
surrounding area.
358

F'i.gure 15. Yoz do Areia Dam-End. of the Hxst phase

figure lb. Segredo Dam - Construction joints and roads in the rockfill
359

Some particular care is usually taken to improve the final trimming


of the upstream face and to minimize the problem of rock pieces falling
down the face of the dam. Further scalping of larger sizes for the
external 1.0 m wide layer reduces the tendency to segregation during the
spreading operation with favorable results. Steel forming was used in
Khao-laemdam, in a labor intense but well succeeded control of the
upstream slope [12].
Although many rockfill dams exist without a special treatment for
the downstream face, it has become a frequent and recommendable practice
to select larger rock blocks to build up a better looking and more
stable surface. The selected blocks are dozed to the theoretical design
line and made to rest on their flat face forming a regular external
layer against which the zone 3C material is compacted, Figure 17.

Figure 17. Specially arranged large blocks result in a regular pattern


for the downstream face of Segredo Dam.

3.2. Concrete works

Construction of the toe slab precedes any concrete work and its early
execution accomplishes two basic objectives: it disciplines the rockfill
construction by clearly defining its upstream limits, and it liberates
the grouting works which can proceed in parallel to the embankment
construction, Figure 18.
360

Figure 18. Plinth at the deeper section - Segredo Dam, transition zone
and rockfill under construction.

The face slab construction is normally carried out by the slipform


method. Triangular starter slabs are initially casted near the plinth,
with temporarily held formwork or by screed concrete, to provide for a
horizontal starting section for the slipform operation of the main body
of the dam.
The system used in Foz do Areia Dam for slipforming the main slab
is illustrative of the method which can evidently accept many alterna-
tives regarding the slipform details and displacement system as well as
the concrete transportation and placement procedures.
The slipforms were metallic structures moved by two hydraulic jacks,
one at each side, supported on rails. The screed length was 1.10 m and
the structure was provided with two platforms; one for the concrete
placing and vibration and the other, at the rear, for surface finishing,
figures 19 and 20.
The concreting operation was preceded by the construction of
concrete pads to support the copper waterstops and the slipform guides.
Positioning of the reinforcement was followed by the erection of the
lateral woodforms, which were used also as supports for the slipform
rails, Figure 20. The slabs were built in alternate pannels. Steel
reinforcing was mounted in place, in such a way as to avoid interrup-
tions in the concreting operation. An average production of 330 ton per
month was attained, with a peak of 500 ton per month, Figure 21.
361

HIDRAULIC JACKS

3400

1600

DIMENSIONS
IN mm

CONCRETE '..----RAIL
SUR FACE

Figure 19. Foz do Areia slipform - lateral view

16000

""
CONCRETE PAD CONCRETE SLAB CONCRETE PAD
LATERAL WOOD FORM FINISHED SLAB

Figure 20. Foz do Arei~ slipform - Section


362

Figure 21. Foz do Areia - Slipforming in alternate pannels - First


stage construction.

The concrete was transported to the site by means of a cableway


and by trucks, and discharged in a transfer hopper located at the top
of the rockfill. Open chutes were used for the transport along the
face of the dam to the concreting section, Figure 22. Care was taken in
vibrating the concrete mass ahead of the form, as to avoid floating
effects in the screed. The simple and economical open chute system
proved to be quite efficient. No segregation occurred in the 8 em slump
well controlled concrete. Concrete curing was carried out by means of
water continuously applied for at least 14 days after placement of
concrete.
Production rate attained a maximum of about 63 m3 /h. Overall
average slip velocity was about 1.50 m/h, including time losses related
to slipform displacement and eventual interruptions. As for the
effective concreting operations, the average slipforming velocity was
2.50 m/h, and the maximum value attained 6.00 m/h. The first stage of
concreting, to EI. 680, was completed in 9 months, and the second stage
took about 13 months. Two slipforms were employed in this work. The
53 slabs of the entire face, with a total area of 140,000 m2 , used a
volume of concrete of 80,000 m3 (including the parapet).
363

Figure 22. Foz do Areia slipform and concrete chute.

The slab volume was about 12.5% larger than the theoretical design
volume, representin~ an average increment of 7 cm in the thickness of
the entire face. Such a small overthickness was due to the special care
taken in the surface finishing of the transition material, and to the
technique of using the forms according to the actual rockfill profile,
instead of the 1.4:1 theoretical slope.

3.3. River handling aspects

The freedom to raise the rockfill to different heights at different


sections, by accepting construction joints and ramps within the body of
the dam, can be used to advantage in planning the construction schedule
and setting the different phases of river diversion.
An upstream section of the dam can be made to act as a cofferdam
even before the placing of the concrete face with significant savings
in the diversion works. In Foz do Areia dam, the construction of a 90m
high first stage dam to act as a cofferdam for a 1:500 yr flood was built
under the protection of a one flood season cofferdam, for a 1:10 yr risk,
Figures 3 and IS.
The semi-pervious transition material and the permeability gradient
resulting from the zoning of the dam control the seapage through the
rockfiJI without endangering its stability.
The experience of R.D. Bailey Dam [21] testifies to the intrinsic
364

safety of the zoned and compacted rockfil I to through flow. Durir;g


construction the water level raised 48 m against the upstream face of
the 95 m high dam essentially completed except for the concrete face.
The upstream transition zone was a 4.3 m wide, minus 10.2 cm crusher-run
sandstone, topped by a 67 cm wide layer of minus 3.8 cm of the same
material, protected by a 7.6 cm thick shotcrete lining to act as a
permanent cofferdam. The shotcrete lining was overtopped by about 6.4m
and seepage through the dam peaked at about 7.1 mS/s, with water
daylighting at the downstream face about 10.0 m above foundation, for a
net head on the dam of approximately 38.4 m. The main body of the dam
was composed of weak shales and thinly-bedded sandstone, resulting if' a
low permeability fill. However, the external rock shells, the cofferdarr.
section,and a 3 to 4.5 m thick bottom drain were from quarried hard
sandstone and provided an efficient drainage which withstood the flood
event without being damaged.
Advancing the embankment construction over the river banks prior
to river diversion is current practice. It was used in a minor way in
Foz do Areia construction and has been fully applied in Segredo dam, as
illustrated in Fi.gure 23. About 1.3 million cubic meters of rockfill
from required excavation at the diversion tunnels were placed on the
right abutment, respecting a 30 m wide strip downstream from the plinth.
Construction of the remaining fill will be in horizontal Jayers after
river closure.

Figure 23. Segredo dam - General view - Plinth, transition in the


canyon and rockfill placed along the right bank prior to diversion,
30 m downstream from the plinth.
365

Such advanced construction can include also an encroachment of the


river section. Underwater dumped rockfill over a rockbottom is being
used in Xingo dam, a 150 m high CFRD in the San Francisco River in
Northeast Brazil, for a 5,000 MW power project, in depths to 20.0 m, to
increase the area for rockfill placement on each abutment prior to river
diversion. The higher deformability of the rockfill has no noticeable
effect on the slab movements if distant enough from the plinth.
As illustrated in Figure 24, hydraulic model tests confirm the
stability of the rockfill slopes to flowing water during floods.

Figure 24. Xingo first phase dam construction encroaches in the river
channel. Flow conditions tested in a 1:100 hydraulic model, CEHPAR
laboratory, Curitiba, Brazil.

In Xingo dam, where closure of the river will involve dealing with
daily peaks of 2,000 to 4,000 m3 /s, and total head differences of 10 to
12 m, three simultaneously built end-dumped dikes are being planned. A
central dike, along the axis of the dam,will greatly facilitate the
closure operation,remaining as part of the permanent dam. By concen-
trating the head difference in the central dike, the closure along the
downstream cofferdam will face a lower head difference avoiding the use
of very large blocks and/or the dragging of the rockfill material,
facilitating the subsequent sealing of the cofferdam.
366

4. MONITORING AND BEHAVIOUR

4.1. Dam movements


Deformations within the rockfill, face slab and joint movements, and
strain distribution in the slabs have been observed in many dams and a
good understanding exists of the general behavior of the CFRD, both
during construction and under load. Measuring the leakage through the
dam is the usual way to check the performance of the waterstop system.
Settlements during construction as measured in Foz do Areia dam by
means of hydraulic cells (Swedish box) indicate the typical vertical
deformation pattern, Figure 25.
INSTRUMENTED 5 ECTIONS

Figure 25. Foz do Areia Dam - Equal settlement curves before reservoir
filling (cm).

Horizontal movements have seldom been measured. Data on Kotmale


dam [15] indicate the horizontal displacements to be negligible during
construction.
The settlement of the central concrete slab under water load is
essentially normal to the face. The perimetric joint at river bottom
practically does not open and the deformation of the slab at the plane
of the face is nominal. Deformations cannot be other than normal. The
general tendency of settlement within the embankment, however, seems to
be steeper, as indicated by the measurements in Kotmale Dam. The state
of compression which results in most of the concrete face suggests also
this tendency for deformations down from normal to the face.
Segredo dam under construction (1990) is being provided with a
system of plate gauges to monitor the horizontal movements normal to the
axis of the dam. The system is similar to the modified British Research
Station type, described by Penman and Charles [16], with one probe for
each plate attached to steel rods externally controlled, and should be
able to further enlighten the question. As for Foz do Areia dam, the
367

basaltic rock in Segredo produces a highly deformable rockfill and is


particularly suitable for the investigation.
The slab deformation at Foz do Areia dam, shown in Figure 26,
i.llustrates some aspects which are typical of CFRD performance.

DEFLECTION (em)
CR 05/80 06/80 07/80 08/80 11/80 10/82 03/84-85
.. .. .. ... ... ..
1-21 2 1.8 27.7 44.0 47.4 50.1 52.5 52.5
7-27 22.8 30.9 55.8 6 1.4 64.8 68. I 70.0
13-33 12.6 2 0.7 5 O. 4 6 1.2 69.2 72.8 77.5
18-38 - i 8. 7 35.6 47.0 56.6 62. I 68.9
\v.L. 702.50 714 .00 73 5.80 739.50 739.00 742 .00 743.60

Figure 26. Foz do Areia slab deformation after reservoir filling.

The deformation normal due to the face is considerably less than


could be expected from the observed construction deformation modulus.
A transversal or water load deformation modulus, computed along the
direction of the water pressure,gives results 1.5 to 3 times greater
than the vertical modulus measured during construction. This has been
observed in all CFRD dams, and can be used as a reference for a rough
estimate of the maximum settlement of the face of the dam.
The face displacements show a sudden increase for the last 20 m of
reservoir filling (June 80 to July 80). The load conditions seem to
overcome a barrier possibly related to the state of stress established
during construction, particularly at the upstream zone of the dam,
368

where confining stresses due to the weight are relatively low. Research
and field measurements are certainly needed for a better understanding
of the deformation pattern, but their results should not change the
practical conclusion that to minimize joint movements and leakage
problems, special care in compaction works is particularly required in
the upstream third of the dam section.
The response to water load in Areia was fast and creep along the
years nominal. The maximum slab deformation reached 70 cm six months
after reservoir filling and increased to about 78 cm after four years,
to become stable thereafter. The behavior of other dams, particularly
in narrower valleys like Cethana in Australia, have shown a lower rate
of deformation immediately after reservoir filling, while creeping,
which seems related to the release of the arch action in the valley
walls, is felt for a longer time. Deformation was 11.8 em seven months
after the reservoir had filled and increased gradually to 14.0 cm after
nine years [17].
Maximum deformations of the face can be nominal in gravel dams like
the 145 m high Salvajina dam (12 cm), or in well graded sound rockfill
such as in Cethana dam (14 cm) or of the order of 80 cm as in Foz do
Areia, without any harmful effect to the slab, basically because the
concrete structure rests freely on the rockfill, the deformation is
gradual, and the ratio deformation to slab-length is always very small.
The resulting joint movements, particularly along the perimetric joint,
are the main concern, because leakage may occur if the waterstop barrier
is damaged.
Vertical joints tend to open near the abutments. In Foz do Areia,
six vertical joints at each side, of a total of 50 vertical joints, have
opened a maximum of 3.0 cm. Total opening for all joints added up to
14.0 em, 1.2 cm in average. In the perimetric joint settlement normally
prevails over the two other components of the joint movement. Maximum
settlement values of 55 mm and 23 mm were measured in Foz do Areia and
Salvajina dams respectively, while openings reached 23 rom and 10 mm, and
tangential movements were 25 rom and 16 rom respectively. The perimetric
joint details were similar, and behaved adequately. Leakage in Foz do
Areia peaked at 240 lis diminishing to a stable flow of about 50 lis
after four years. Salvajina dam reported a maximum nominal leakage of
60 lis with a clear tendency to decrease also.
As the membrane follows the rockfil1 movement, it is dragged
towards the center of the dam and placed under compression, except for
the upper zone and along the plinth, as shown in Figure 27.
Slab strains have been measured at mid height of the slab and have
not indicated torsion or flexion stresses possible to occur at the
proximity of steep valley walls. In Segredo dam, strain-gages are being
installed at both sides of the central reinforcing mesh in an attempt
to detect such anomalies in the strain distribution picture.

4.2. Performance under seismic load

As a free draining structure the CFRD dam on rock foundation is


essentially safe against earthquakes. No pore pressures may develop
and no risk of liquefaction exists.
369

B
Figure 27. Equal strain curves (x 10- 6 ) at the concrete face - Foz do
Areia Dam [14].
A - In the slope direction B - In the horizontal direction

The performance of Cogoti Dam, an old dumped CFRD in Chile (1938),


under seismic loads due to a 7.9 magnitude 1943 earthquake, with
epicenter located 89 km away, with ground acceleration estimated to have
reached 0.19 g, is the best existing evidence on effects to be expected
in modern compacted CFRDs [18]. The embankment settled 40 cm or 0.45%
of its 84.0 m height due to the shaking. The seismic load apparently
accelerated the settlement of the fill which could be expected in the
long-term, as illustrated in Figure 28. Three subsequent earthquakes,
with magnitude of 7.1 to 7.7, and ground acceleration of 0.03 to 0.05 g
produced no noti.ceable additional effects.
No seismic damage to the concrete face slab was reported, except
for soule minor crest damages. Exposed joints are intact, but leakage
through the dam has increased along the years and may be related to
cracks and joint movements induced by the relatively deformable dumped
rockfill. There was no question of slope stability either upstream
where the water effect is benefitial or downstream.
For modern compacted CFRD dams the deformation due to shaking
should be considerably less. Sherard [19] has estimated it to be about
0.1% of the dam height for a dam of the same size and under identical
seismic conditions. Cracks in the concrete face and increased leakage
would be of no concern for the stability of the dam.
370

5.0
(1524)
L I, 943 E~RTHQlJA~E J. I I 971 ' EARTHciuAKE 1965 bRTHQUAKE

I
I 965 EARTHQUAKE
<f)

ffi 4.0
f- (1219 )
W

-- -
~
:::;
~ 3.0 ~
r-
(914 )
f-
W
W
I"'V ~ .....
..... ---
u.. .....
!,'; 2.0
..... .--
(610) ,/
I-
Z ,/
w
~
J"
~
1.0
(305)
~

f
<f)

o
'"0>
OJ OJ OJ
'"0> OJ
'"
0>
<Xl
0>

YEARS

Figure 28. Crest settlement curve - Cogoti Dam [18]

An evaluation of the conditions to be expected can be roughly


assessed from the earthquake severety index (ESI) as defined by Bureau
and others [20]. Based on an average relationship between earthquake
duration D, in seconds, and magnitude M,

D = 7 (M - 4.5)1.5

the product

ESI = PGA (M - 4,5)3

where PGA is the peak ground acceleration, was defined as the earthquake
severity index, to which earthquake induced deformations in rockfill
dams were correlated as shown in Figure 29.
The observed values in Figure 29 refer either to a dumped CFRD as
Cogoti or to earth core rockfill dams, such as La Villita and Infernillo
dams, more deformable than compacted CFRDs. Although such correlation
curve as pointed out by its author should not be used for predicting
settlements for new or existing dams, they support Sherard and Barry
Cooke's conclusion that "for the great majority of sites which may be
very strongly shaken such as near the epicenter of a Magnitude 7.5
quake, or at sites with calculated Earthquake Severity Index in the
general range of 10-15, the same CFRD design can be used as in non-
seismic areas. For these sites, all present experience with dam
behavior and the overall results of current dynamic calculations give
confidence that the worst earthquake-induced crest settlement will be
substantially less than 1% of the dam height. A sudden crest
settlement of 1% of the dam height will not treaten the safety of a
modern CFRD" [6].
371

Observed /
Computed

~ La Villito Dam
\l El lnfiernillo Dom
ESTIMATED
oIO.O-~----~A~V~EA~G~E~RE~L~A~T~IO~N~SH~I~P~~--~----t
Q'!.
I-
Z
W
:::>
w
--'
I-
hi
<f)

<i I.O--lf-------+--_=_
u
i=
a::
I~I
w OBSERVED
> AVERAGE RELATIONSHIP
w
>
~
--'
w
a::
O.I-+-----e..---,,,L--+------+------t

ROCKFI LL DAMS

0.1 1.0 10.0 100.0


EARTHQUAKE SEVERITY INDEX (ESi)

Figure 29. Relationship between crest settlement and earthquake


severity index [20].

In cases that even stronger earthquakes could occur, careful


analysis should be carried out. Ample and conservative freeboard is
probably the most appropriate provision for safety against earthquake
effects.

l<EFERENCES

1. Wegmann, Edward (1908), The Design and Construction of Dams,


John Wiley & Sons, New York.
2. Barry Cooke, J. and Arthur G. Strassburger (1988), "Rockfill Dams",
Section 6, in Development of Dam Engineering in the United
States, USCOLD, Ed. by Eric B. Kollgard and Wallace L. Chadwick,
Pergamon Press, N. York, 885-897.
3. Leps, Thomas M. (1988) "Rockfill Dam Design and Analysis", Chap.12,
Advanced Dam Engineering for Design, Construction, and
Rehabilitation, Ed. ~obert B. Jansen, Van Nostrand Reinhold,
New York, 368-387.
372

4. Steele, I.C. and J. B. Cooke (1969) "Concrete Face Rockfill Dams",


in Handbook of Applied Hydraulics, Davis and Sorensen, Third Ed.,
McGraw Hill Book Co., 19.1-19.16.
5. Barry Cooke, J. (1984) "Progress in Rockfill Dams", 18 th )'erzaghi
Lecture, ASCE, Journal of Geotechnical Engineering, Vol. 110, No.
10, 1383-1413.
6. Barry Cooke and James 1. Sherard (1987) "Concrete Face Rockfill
Dam", Assessment and Design, ASCE, Journal of Geotechnical
Engineering, Vol. 113, No. 10, 1096-1132.
7. Steele, I.C. and J.B. Cooke (1960) "Rockfill Dams: Salt Springs and
Lower Bear River Concrete Face Dams" Symposium on Rockfill Dams,
Trans. ASCE Vol. 125, Part II, 74-159.
8. Leps, Thomas ,M. (1970) "Review of Shearing Strength of Rockfill",
ASCE, Journal of the Soil Mechanics and Foundations Division,
Vol. 96, No. SM4, 1159-1170.
9. Sierra, Jesus M., Carlos A. Ramirez, and Jorge E. Hacelas (1985)
"Design Features of Salvajina Dam", ASCE, Geotechnical Engineering
Division, Detroit Symposium on Concrete Face Rockfill Dams,
266-285.
10. Pinto, Nelson L. de S., Bayardo Materon and Pedro Lagos Marques
Filho (1982) "Design and Performance of Foz do Areia Concrete
Membrane as Related to Basalt Properties", Q.55, R.51, 14th ICOLD
Congress, Rio de Janeiro, 873-906.
11. Pinto, Nelson L. de S. e Rui T. Mori (1988) "A New Concept of a
Perimetric Joint for Concrete Face Rockfill Dams", Q.61 R.2, 16 th
ICOLD Congress, San Francisco, 35-51.
12. Watakeekul, Som Kuan, Gordong Roberts, Andrew J. Coles (1985)
"Khao Laem - A Concrete Face Rockfill Dam on Karst", ASCE
Geotechnical Engineering Division, Detroit Symposiu~ on Concrete
Face Rockfill Dams, 336-361
13. Pinto, Nelson L. de S., Pedro L. Marques Filho, Edilberto Maurer
(1985) "Foz do Areia Dam - Design, Construction and Behaviour",
ASCE, Geotechnical Engineering Division, Detroit Symposium on
Concrete Face Rockfill Dams, 173-191.
14. Marques Filho, Pedro L., Edilberto Maurer, Nelson B. Toniatti (1985)
"Deformation Characteristics of Foz do Areia Dam Concrete Face
Rockfill Dam, as Revealed by a Simple Instrumentation System",
Q.56-R.21, 15 th ICOLD Congress, Lausanne, 417-450.
15. Gosschalk, Edward M. and A.N.S. Kulasinghe (1985) "Kotmale Danl and
Observations on CFRD", ASCE, Geotechnical Engineering Division,
Detroit Symposium on Concrete Face Rockfill Dams, 379-395.
16. Penman, Arthur, J.A. Charles (1982) "An Improved Horizontal Plate
Gauge", Building Research Establishment, Technical Notes,
pp.278-282.
17. Fitzpatrick, M.D., T.B. Liggins, R. H. W. Barnett (1982) "Ten Years
Surveillance of Cethana Dam", Q.52, R-Sl, 14th ICOLD Congress,
Rio de Janeiro, 847-866.
18. Arrau, LuIs, Ismae] Jbarra, Guillermo Noguera (1985) "Performance
of Cogoti Dam under Seismic Loading", ASCE, Geotechnical
Engineering Division, Detroit Symposium on Concrete Face Rockfill
Dams, 1-13.
373

19. Sherard, James L. (1987) Discussion to "Seismic Analysis of Concrete


Face Rockfi11 Dams", ASCE, Journal of Geotechnical Fl1gheering,
Vol. 113, No. 10, 1252-1254.
20. Bureau, Gilles et a1 (1985) "Seismic Analysis of Concrete Face
Rockfi11 Dams", ASCE, Geotechnical Engineerh,g Division, Detroit
Symposium on Concrete Face Rockfi11 Dams, 479-508.
21. Beene, Ralph R.W., Edward C. Pritchett (1985) "The R.D. Bailey Dam-
A Concrete-Fnced, Earth-Rockfi11", ASCE, Geotechnical Engineering
Division, Detroit Symposium on Concrete Face Rockfi11 Dams,
163-172.
22. ICOLD, "Rockfi11 Dams with Concrete Facing - State of the Art",
Bulletin 70, 1989.
CHAPTER 14
STATIC BEHAVIOUR OF EARTH - ROCKFILL DAM S
E. MARANHA lias NEVES

I. Introduction

Fill dams (or embankment dams) were the first dams built by
man, they are the most nllmerous and the most frequently chosen
when it is a question of constructing a new dam. According to
Penman (1986), in 1985 statistics showed that over 80% of the
large dams (higher than 15 m) constructed in the world were
earth dams. Nowadays that figure is certainly higher. And it
must be added that the highest uams in the world are also fill
dams.
A type of fill dam that has become very c ommon is t h e
rockfill dam with a central core. This ,;olLltj.on is due to a
desire for optimization in th~ us e of available natural
materials by placing them in the best zones of the dam body.
I n the forties, there began to be c0nstruc ted works of this
type , a n d their conception has owed a lot to the great
progress made in soil mechanics, whi c h has been concerned
mainly with shear strength and filter s . In the last 20 years,
and as result of problems arising as regards cracking i n
cores, especially hydraulic fracturing, important studies have
been carried out on the conception of these works, and in them
have been incorporated some of the more important advances in
soil mechanics, namely regarding stress-strain - strength rela-
tions, seepage in foundation ground and in fills as well as
filtering and drainage phenomena .
Nowadays earth - rockfill dams, both with central core and
with sloping core, have reached un precedent heights (300 m in
the Nurek Dam and 335 in the Rogun Dam, both in the USSR), and
constitute one of the most frequent solutions in the domain of
fill dams. It seems that stabili ty problems do not arise and
capacity for mastering the questions connected with defor-
mability has been substantially improved si n ce the sixties ,
with introduction of the tec hnology of compacting rockfill by
using vibratory rollers.

'"
E. Moronho dos N~I'~S (~d.J, Ad,'onus in Rod.fil/ SlruClltr~s. 375-447.
e 1991 Kluwf'r Acodf'mic Publishers.
376

Nevertheless, ensuring the safety of a dam of this type is


not confined to guaranteeing the absence (or very low proba-
bility of occurrence) of failures by sliding. In fact, any
defect in the filtering arrangement of the core may lead to a
localized phenomenon of internal erosion (piping), of a
progressi ve kind, or to loss by suffusion of appreciable
quantities of material from the core that go to fill the voids
of the rockfill mass situated downstream. Piping in the core
may also be set off by the occurrence of cracking of that
watertight component. This means that this study of the
equilibria leads to study of the compatibility of the deforma-
tions. It must, however, be emphasised that the present
tendency towards the use of rockfills consisting of low
quality rock materials (low resistance rocks, for example)
makes it pertinent to analyse the limit equilibria.
Analysis of the behaviour of these structures in the light
of the various actions and for different design situations has
proved to be a matter of considerable complexity, but progress
in the mechanical characterization of particulate media in
general, and development of methods for quantifying structural
behaviour make it possible to regard with optimism the problem
of safety of these eminently geotechnical works. The same can
be said as regards serviceability.
Lastly, there is no doubt that validation of the procedures
that have been set out above depends on the monitoring of
dams, and that monitoring constitutes a source of information
which is of vital importance.
Some authors have made appreciations on the structural
behaviour of earth-rockfill dams, based on the results of
observation (for example, Wilson, 1973; Justo Alpanes et al.,
1985; Leps, 1988, etc.). Sometimes the criteria used in
choosing the dams analysed are not made explicit.
The next section will refer to a series of earth-rockfill
dams for which monitoring results are available.
This set of dams was selected from among a very numerous
series, with the aim of obtaining a wide spectrum information
that would be sufficiently representative. Through this set of
dams an effort will be made in section 3 to document some
conclusions on existing predominant orientations in the
conception of earth-rockfill dams.
In section 4 contribution of mathematical modelling for
structural conception of earth-rockfill dams is highlighted
and in section 5 matters dealing with safety analysis are
tackled. Finally in section 6 some broad conclusions are
pointed out.
377

2. Geometrical Physical and Mechanical Data of a Series of Earth - Rockfill Dams

Before analysing the available information, it is necessary to


explain whether the series presented is significant and which
are the parameters that really require attention.
As regards the first point, it was considered that a set of
dams would be significant when there was as much variation as
possible in the characteristics of the dam that rould most
reflect or mark its behaviour, e.g. the geometrical elements
(namely height, location and thickness of the core, width and
inclination of the slopes), the kind of materials used in
construction and the specifications for their placement (in
particular quantity of water added and compactness).
As regards the parameters to be taken account of, these will
first of all be those measured by the apparatus installed,
i.e. total, effective and pore water pressures, deformations
and seepage flows. But another type of information that may be
very important is that resulting from visual observation. This
concerns all types of incidents (equivalent or not to service-
abili ty limit states <1J), principally concentrated deforma-
tions, cracks and springs.
Another aspect to be pointed out is that most of the dams
considered belong to the last generation, i.e. they incorpo-
rate to a greater or lesser degree some of the most recent
improvements as regards conception, execution and monitoring.
Older dams are included only when they have particular
features considered to be useful in the study concerned.
Lastly, attention is called to the enormous geographical
dispersion, intended thus to introduce geological and meteoro-
logical conditioning factors and, above all, local experience.
Table 1 gives in chronological order the series of dams
analysed, showing the most significant characteristics of each
work.
Figs. 1 to 12 show the characteristic cross sections of the
dams of Table 1. This is, in fact, complementary information
that is important when the structural behaviour of these
structures is discussed.
The section that follows will deal with aspects that are
envisaged as fundamental as regards the structural behaviour
of these structures. The exemplification will make use prima-
rily of information relating to this series of dams. That is
to say, it is not intended to draw conclusions from a compara-
tive analysis of the behaviour of that series. The aim is
rather to use it as support for considerations based on a far
vast range of information on the hundreds of high earth-
rockfill dams existing throughout the world.

(1) Reference will be made later, in more detail, to the ser-


viceability limit states when dealing with the question of
dam safety. the former can be planned separately.
378

A note, also, on the foundations, which of course are an


important part of the dam. It is logical that they should be
analysed in conjunction with the body of the dam. But except
in cases in which an earth-rockfill dam is founded on soil,
the contrast between the deformabilities of the fills and of
the foundation ground is so marked that from this point of
view the body of the dam can be studied separately. The same
may be said, as regards permeability, but the situation is
symmetrical to the previous one: the permeability of the rocky
foundation (a more or less fissured medium) is normally higher
than that of the core, so that eventual treatment of the
former can be planned separately.

Table 1. Main characteristics of a set of earth-


rockfill dams.

r" 1 iNFtERNiLLO
DAM CCXJNTRY

MEXICO
COMPLETED
I'
1963
L/H
1m)

344/148=2.3
CORE
POSITION

VERT I CAL
SLENDERNESS
INDEX
(B/Hw)

0.26
LONG J TUD] NAL
AX IS

STRAIGHT

2 AKOSOMBO GHANA 1965 622!11~=6.0 VERTICAL 0.35 STRAIGHT

3 SCAMMONDEN U, K. 1969 622170=8.9 SLOPING 0.44 STRAIGHT

4 TALBINGO AUSTRALIA 1970 729/162"4.5 SLOPING 0.63 STRAIGHT

5 ANGUSTURA MEXICO 1974 300/145"2.1 NEAR VERTICAL 0.34 STRAIGHT

6 SVARTEVANN NOR\JA Y 1976 400/129"3.1 SLOPING 0.22 CURVED

7 CttICQASEN MEXICO 1980 306/261"1.2 VERTICAL 0.38 STRA j GHT

8 EMBORCA~AO BRAZI L 1981 1607/158=10,2 SLOPING 0.55 STRAIGHT

9 MONASAVU FIJI 1982 460/85 .. 10.2 SLOPING 0.48 STRAIGHT

10 LG4 CANADA 1983 800/125=6.4 VERT J CAL 0.57 STRA: GHT

11 BEL I CHE PORTUGAL 1985 560/54"10 VERT leAL 0.57 CURVED

',2 CARACOl MEX ICO 1985 360/126"2.9 VERT! CAL 0.56 STRAIGHT

References: - MarsaL, R.J. and l. Ramirez Arellano (1967); Gonzalpz Valenc1a, F.


and M.A. Gusman Martmez (1985)

- Wi [son, S.D. (1973)

Charles, J.A. (1973); Perwnan, A.D.M. and J. Charles (1973); PerYTlan.


A.D.M. (1986)

- Adikari, G.S.N. and A.k. Parkm (1982)

GonzaLez VaLenc1a and F. SaLdana G{)IJI;'>S (1985)

DiBiagio, E. et at. (1982); I(Ja('rnsLl, 8. et at. (1982)

7 GonzaLez Valencia, F. and L. Aguirre SorIa (1985)

Mello, V. de (1983); MeLLo, V. eX> (1984) Parra, p.e. (1985)

9 - Knight, O.J. et al. (1983)

10 Mc ConneL, A.D. et al. (1982); Verma, N.S. et aL. (1985)

11 LNEC (1966a); NayLor, D.J. et al. (1986); LNfC (1989)

12 - C~s Pina, J.M. and M.A. Guzman Martmez (1985)


379

Table 1 (Cont.)

ROCKFllL CORl
NO DA" SEISMIC
MATERIAL COMPACT ION MATERIAL WATER FOUNDATION ACT IONS
TECHNIQUE CONH.NT

, iNFIERNILLO SILICIOUS VIBRATORY ROLLER CLAY +2/0 TO +4% SILICIOUS


CONGLeJoIERATE & 08 OUTER SHELL CONGLOMERATE YES

2 AKASOMBO QUARTZITE WITH VI BRATORY ROLLER HIGHLY o TO +2% ROCK


SHALE PLASTIC CLAY NO

3 SCAMMONDEN SANDSTONE VIBRATORY ROLLER CLAY \.j! TH +rt, TO +20% SANDSTONE SHALE
SHALES NO

4 TALBINGO RHYOL ITE VIBRATORY ROLLER CLAY +0.3% RHYOLITE; TUFF


NO

5 ANGOSTURA LIMESTONE DUMPING &. CLAY -2% TO 0 ([HF) I_IMESTONE


VI BRATORY ROLLER o TO +?Y, (SUP) YES

6 SVARTEVANN GRANITIC GNEISS VIBRATOR)' ROLLER TILL o TO +2% GRAN'T Ie GME I 55


ND

7 CHICOASEN LIMESTONES iJITH DUMPING & CLAYEY SAND -D.B}: TO [I LIMESTONE


SHALES VIBRATORY ROLLER YES

8 EMBORCAt;AO GRAN 1 TE GME I SS VIBRATORY ROLLER PLASTIC CLAY , .S% TO 0 ROCK


NO

9 MONASAW MONZONITE VIBRATOR'!' ROLLER HIGIIL Y -+-20% MC~ZONITE SANDSTONE


PLASTIC CLAY

,0 LG4 GRANITE; GNEISS VIBRATORY ROLLER Tt LL 1% TO -+-?'Y.- GRANITE; GNEISS


NO

" BELICHE SCHIST; GREYIJACKE VIBRATORY ROLLER CLAY - 1% TO +2% ALLUVIA; SCHIST;
GRFYWACKE YES

12 CARACOl SANDSTONE VIBRATORY ROLLER CLAY o TO -+-1% ALLUVIA; SANDSTONE;


SILTSTONE; SHALES YES

Problems with concentrated flows through the core can only


arise if there is piping or cracking, and these phenomena are
the object of our attention.
For these reasons, and owing to the need to limit the scope
of this work, the foundations, or seepage through them, will
only be referred to en passant.
380

-l2~_O.OO

0 SOm
t===t
8 -IMPERVIOUS CORE (i)-DUMPED ROCKFILl

o -FILTERS 0- RIPRAP

0- TRANSITION ZONES 0- CUT-OFF WALLS (SECANT PILES J

o -COMPACTED ROCKFILL CD-ALLUVIUM

Figure 1. Cross section of Infiernillo Dam.

SOm
I

(0 _CLAY CORE (3) - ROCKFILL l COMPACTED)

o -FILTERS ({;, - DUMPED OR COMPACTED ROCKFILl

Figure 2. Cross section of Akosombo Dam.


3RI

" ~6'i50
--"'---

"
- 1,

- CLAY CORE

- ROCKFILL

Figure 3. Cross sect.icr. of Scammondeli Dam .

. 5~2 0.9
_ _ ---=t::::.

-~~ -

2B
3C

o
=
SOm

0)-CLAY CORE @.@ - SOUND ROCKFllL

@-FILTERS B- WEATHERED ROCKFILL

@- TRANSITIONS @- RIPRAP

Figure 4. Cross section of Talbingo Dam.


3X2

o
=
100m
(~) -CLAYE"Y SOIL "
i I. i-OUMPED
~
ROCKFILL

(~) - COMPACTED SAND AND GRAVEL (,,-RIP


~
RAP 0- COFFERDAM

(~) -COMPACTED ROCKFILL G- ALLUVIUM o -IMPERVIOUS WALL

Figure 5. Cross sec-cion of Angostura Dalli.

90350

o SOm
L===:t:===---t

8- MORAINE CORE ~'- TRANSITION ROCKFILL 0- RIPRAP

0- SANDY GRAVEL FILTER -ROCKFILL

Figure 6. Cross section of Svartevann Dam.


o
'8 -
100m
- CLAY CORE C:OMPACTED ROCKFILL
Cc===t

,G~ - CLAY MATERIAL I WET SIDE) 8- UNIFORM ROCKFILL 115'" d n ..: 25 em) 0- COFFERDAM

-FILTERS DUMPED RDCKFILL (~-IMPERVIOUS WALL

- TRANSITION ZONES 0- ALLUviUM (~) -RIPRAP

Figure 7. Cross section of Chicoasen Dam.

529_00
~-

o 'SOm
t==t==t

8-CLAY CORE 8-COMPACTED ROCKF1LL ILAYERS. 06ml

@-COMPACTEO ROCKFILL ILAYERS, 1.2 ml -IMPERviOUS RAMOOM MATER IAL

@,-COMPACTED ROCKFILL (LAYERS. 0 9 m) 8-LARGE BLOCKS

0- FILTERS

Figure 8. Cross section of Emborca<;:ao Dam.


384

'"
"

50 m

(11 CLAY CORE

FILTERS,DRAINS AND TRANSITIONS

Figure 9. Cross section of Monasavu Dam.

3. 3A ,-2

Sam
I

@-SA.NOY GRAVEL ID mox ,,7.5cml @- ROCKF ILL I Dmax ~ 2m I

~I-SANDY GRAVEL ! 0max= 30 em! (4)- RIPRAP

Figure 10. Cross sect,ion of LG4 Dam.


385

(0 - CLAY CORE 0-
ROCKF1LL

0~FILTER AND ORAIN 0- ALLUVIUM

o -ROCKf"ILL (WEATHERED 1 8- DIAPHRAGM WALL

Figure 11. Cross section of Beliche Darn.

o 50 m
t====

(2)-CLAY CORE G)-OUMPED ROCKFILL 0-RIPRAP ~LARGE BLOCKS)

0-FILTER (i)-ALLUVIUM

G:::,- TRANSITION (i)-WASTE MATERIAL FROM ROCK BLASTING OPERATIONS

~- COMPACTED ROCt<FllL ~}-CUT-OFF WALL

Figure 12. Cross section of Caracol Dam.


386

3. Structural Behaviour and Experience

The considerations that will now be offered on the structural


behaviour of earth-rockfill dams are based on available
experience, and it is common place to stress the decisive part
played by monitoring in gaining and distilling that experi-
ence. Later on, there will be the opportunity to confront this
means with that of mathematical modelling.
Although it is extremely interesting and useful, it does not
seem appropriate, in this kind of work, to give an account of
the evolution during the last 40-50 years of the structural
conceptions of earth-rockfill dams. Only a brief reference
will be made to the fact that the advent of the technique of
compaction of rockfills by vibration has been far less
determinant in perfecting solutions in earth-rock-fill dams
than in those of concrete faced rockfill dams (CFRD). This is
due to the fact that in the latter the problem of
deformability of the rockfills - above all the deformations
due to creep - conditions the possibility of assuring the
watertightness of the concrete membranes. In fact, principally
for dams with high height, only a high compactness of the
fills, which is impossible to achieve in a practical and
economical way without resorting to vibrator rollers of an
appreciable static weight, can ensure proper (watertight)
behaviour of the concrete faces. As proof of this point of
view, there are the well known examples of the dams of Salt
springs (1930) and Paradela (1958), with heights of 100 m and
110 m, respectively, in which serviceability was really
affected (significant passage of water through the concrete
face) .
Accordingly, although relative deformability is always
concerned in both types of dam (concrete slab/rockfill in one
case, core/rockfill in the other), in absolute values it is
essential that high moduli of deformability should be required
for the rockfills which bear the concrete faces of dams with
a height of over 70-90 m.
It may only be mentioned that while before 1940 interesting
CFRD were constructed, using dumped rockfill and upstream
reinforced concrete face, it was only from that time that
there appear rockfill dams with core that are well exempli-
fied, such at Nantahala Dam (USA), in which use is made of a
steep and narrow core that is protected on both sides by
filters with various thicknesses and grain sizes (see Fig.
13) .
The conception of this dam, as well as of a set of six other
analogous dams constructed at the same time in the united
states, proved to be very good and was influenced by the soil
mechanics concepts, namely as regards filters and shear
strength. It was a radical breakaway from a previous practice
characterized by the use of clay fills upstream, supported on
387

dumped rockfills downstream. There were no filters between the


two zones, which was responsible for several accidents.

=
o 30 m

(2) - CLAY CORE @- QUARRY RUN ROCK I SMALL SIZES)

CD-FILTERS AND TRANSITIONS @- QUARRY RUN ROCK (INTERMEDIATE SIZES)

0- aUARRY RUN ROCK @-aUARRY RUN ROCK (LARGER SIZES)

Figure 13. Cross section of Nantahala Oam.

As the result of good experience with dams of the Nantahala


type, similar structures were built in several countries, but
of special note is Brownlee Dam (USA), constructed in 1958,
owing to its having been founded on an alluvial deposit of
high relative density consisting of sand and gravel with
boulders (Fig. 14).

'561,00

5B 5C
\
~,~"c_~-_~_c __ e,~_-,-c,-,~--~

50 m
1:====11

.~
'~) - CLAY CORE - SMALL ROCK FILTER ZONE -ROCKFllL (LARGER SIZES)

,8, - FINE FILTER ZONE - ROCKFlll (SMALLER SIZES) ~) - ALLUVIUM (SAND,GRAVEl,AND BOULDERS)

G COARSE FILTER ZONE -ROCKF1LL (INTERMEDIATE SIZES) 0-0UMPED ROCKFILl ID m1n 12m)

Figure 14. Cross section of Brownlee Darn.


388

In the following some considerations will be made about the


influence of the construction materials, placement techniques
and conceptions of the structure on structural behaviour of
earth-rockfill dams.

3.1. CONSTRUCTION MATERIALS

There are hardly any limitations as regards the materials for


the stabilizing shells and for the core. In fact, as far as
the rockfills are concerned, rock materials of all types have
been used, and this is not surprising as long as the rock has
good indices for mechanical strength and weathering. What it
is most interesting to point out is the increasing use of
weathered rocks or rocks of low mechanical strength, as is the
case with Beliche Dam (schists and weathered greywackes, the
latter with unconfined compressive strength of 60 kN/m2) and
Scaromonden Dam (sandstones with unconfined compressive
strength of 70 kN/m 2 ).
As regards the cores, the picture is similar. Soils may have
a whole range of characteristics, from non-plastic materials
to highly plastic materials. An example of the latter is to be
seen in the case of Monasavu Dam, where in the core clay was
used with a liquidity limit and plasticity index of about 110
and 50%, respectively. Even extremely adverse conditions as
regards the natural water content of the soil for the core
(very high values) have not been an unsurmountable obstacle,
and suffice to mention the Swedish experience with "wet
compaction" and the case of Monasavu Dam, mentioned above, in
which the water content in compaction of the core material was
76%!
Also the conventional requirement of plastic soils for the
core no longer had any sense, for reasons that will be
presented latter on. It is necessary to have a soil with a
coefficient of permeability which is compatible with the
watertightness needed for the core. It happens that low
permeability is not a property that is confined to plastic
soils or those which contain a significant fraction of
particles with clay dimension. Well known, at least, are the
applications of glacial alluvia in cores, as is the case with
many dams in Sweden, Norway, Canada, etc. In fact, these are
soils that have hardly any fine fraction (dimension less than
0,074 rom), obviously not plastic, but with granulometric
characteristics that are close to Fuller's curve and therefore
give them a reduced permeability (about 10- 6 cm/s, like a
clayey soil). As an extreme situation, mention may be made of
storvatn Dam (Arnevik et al., 1988), where there was analysed
the hypothesis of processing material for the core from a
quarry, and making use of technologies identical to those used
for making materials for filters. A material with a grain size
similar to that one of a moraine would be obtained. Although
it was not the solution that was chosen, it was considered a
feasible procedure.
389

3.2. PLACEMENT TECHNIQUES

As-regards core materials, there is nothing special to note.


In fact, the more generalized specific compaction energy is
that one corresponding to the normal Proctor test, and the
choice of water content for placement that is close to the
optimum or slightly on the wet side is current practice, even
though there are exceptions, as in the case, for example, with
Infiernillo and Monasavu Dams.
As regards the rockfills of the shells, they are more often
compacted in layers with a thickness between 0.6 and 1.5 m.
Placement of these materials by layers makes use of the
technique of laying by dumping with consequent segregation,
and this gives a stratification to the shells (Cooke, 1984).
Normally vibratory compaction is used, and in the case of
certain types of rock, water must be added to the fills in an
appreciable amount. It is desirable that the maximum dimension
of the blocks should not surpass the thickness of the layer.

3.3. STRUCTURAL CONCEPTION

Except for aspects concerning details - which does not make


them less important - such as crest details, protection of the
upstream slope, connection of the core to the foundation and
to the hydraulic structures, etc, conception aspects are
basically connected with core slenderness, inclination of the
slopes, position of the core, relationship between deforma-
bilities of the different zones of the dam, filters upstream
and downstream of the core, shape of the valley and configu-
ration in plan of the dam.
An effort will now be made to dealt with each of those
aspects, and to explain and justify what are considered to be
the most up-to-date and suitable views on this matter.

3.3.1. Core Slenderness. The examples presented in Table 1


show the great variety of situations. Expressing slenderness
by means of the index B/H", in which B represents the maximum
core thickness and Hw the difference between full storage
level in the reservoir and the level of core foundation, it is
seen from Fig. 15 that slenderness index, which varies from
0.25 to 0.65, cannot be correlated with the height of the dam,
with the nature of the soil or rockfills, with the position of
the core and with the inclinations of the slopes. But in fact
the values most frequently adopted are higher than 0.33 (more
or less one third of the maximum head water on the core
foundation, expressed in height of the water). The reasons for
higher (or lower) values are mainly bound up with the rela-
tionship between the deformabilities of the core and the
rockfill shells, and greater or lesser reliability of the
filters situated at the downstream slope of the core. For
reasons that will be given later, these filters are known as
390

"critical filters" (Sherard et al. I 1984b). If seismic actions


of a significant kind are likely to occur l an aspect that is
not dealt with in this work they may lead to the adoption of
I

less slender cores.

z!!\i
Him}
f

.
Hw
280
7 }
~B-;

- vertical core
0- sloping core

200

8 4
CD CD
5
6 CD
12
CD -10
120

9
3 CD

40
CD
11
.
O~------~------~------r-----~
0.2 0.4 0.6 0.8
SI Hw

Figure 15. Slenderness index (B/Hw) versus dam height (H)


Plotted numbers refer to dams presented in Table 1.

certain positionings of the core may constitute a reason for


making it very slender. An example is Brownlee Dam (Fig. 14)1
where the index of slenderness is practically equal to 0.16
but the corel precisely on account of its position (inclina-
tion of 1.4H:1V)1 from a stability point of view conditions
the upstream slope l which assumes an unusually gentle inclina-
tion (2. 6H : 1 V) .
3.3.2. Inclination of the Dam Slopes. Unlike what occurs in
earth dams I where the inclination of the slopes shows a wide
range of variations this does not occur with earth-rockfill
I

dams. Fig. 16 represents the limits of variation for the


inclination of the slopes of the dams shown in Table 1. Except
in the case of the upstream face of Scammonden Dam (in which
the reduced inclination is due to conditioning factors in the
position of the core owing to the installation of a highway on
the dam crest l whose width is thus approximately 57 m)1 it is
found that the inclination of the upstream faces varies
between 2H:1V and 1.6H:1V and of the downstream faces between
2H:1V and 1.4H:1V.
In view of the wide range of characteristics of the dams
being ana1ysed it may be concluded that the matter of the
l

inclination of the slopes does not determine their structural


391

(or safety) behaviour. In fact, most of the rockfills using


low strength rock in dams always behaved well from the point
of view of shear strength and permeability (Justo Alpanes et
al., 1985). The same cannot be said as regards deformability,
but this property has a far less determinant relationship to
the inclination of the slopes than has shear strength.

-- ---,
, 1.6
--2

1.4
,

- - SCAMMONDEN

Figure 16. Range of inclinations of the outer slopes of the


dams of Table 1.

3.3.3. position of the Core. The first earth-rockfill dams


with core are characterized by a marked inclination and
reduced thickness of this component (see Figs. 13 and 14, for
example) or by using a very thick vertical core, as is the
case of Cherry Valley Dam, built in 1955 in the USA, with a
height of 96 m (Lloyd et al., 1958). Its typical cross section
is shown in Fig. 17. In both cases, the rockfills were simply
dumped. Between 1955 and 1965 a transition took place to
compacted rockfills, with earth-rockfill dams reaching very
great heights. In fact, the introduction of a new construction
material - compacted rockfill - made possible the use of low-
strength rocks. When compacted in fairly thin layers, with
water added, this rock provides a material with a dry unit
weight that is very high, and thus a fill that is very
resistant and - what is very important - with low compress-
ibility. In the case of good quality rocks, on the other hand,
the rockfill has less compactness and' its strength is due
mainly to the high interlocking between blocks.
From 1965 until nowadays, cores have been either vertical or
sloping, with variable slenderness. Fig. 18 tries to translate
that variation for the dams referred to in Table 1.
The cores no longer assume the position close to the upstream
face that was pointed out in 3.3.2.
As regards the reasons that may lead to choosing a vertical
or a sloping core, it is found that they do not point in a
definitive manner to either solution. There are, to be sure,
392

reasons that may lead to the choice of the position of the


core, but they have nothing to do with structural behaviour;
there might be, for example, a lack of available soil, which
would favour the vertical solution. Sherard et al. (1963)
state that it is sometimes argued that structures with a
sloping core may be less subject to cracking due to differen-
tial settlement. But the same authors question the soundness
of that argument, though they say that they have no well based
opinion on the matter. The referred authors expressed also
doubts about an eventual better behaviour of sloping cores
from the point of view of cracking and other deteriorations
due to seismic action. They only affirm that rockfill dams
with a sloping core are basically safer when subjected to
earthquakes owing to the existence of a stable rockfill shell
downstream of the core, and not because the core is less
likely to crack.

,r:;2

8 -CLAY CORE (!) - COARSE F!LTER

o -FINE FILTER ( ~) - DUMPED QUARRY RUN ROCK

Figure 17. Cross section of Cherry Valley Dam.

Penman and Charles (1973) put forward some very interesting


view on this theme, but confined themselves to the case of
Scammonden and Llyn Brianne Dams, and "their main aim was a
relationship between the position of the respective cores
(compacted on the wet side) and the horizontal deformations in
the rockfill zone after filling.
V. de Mello (1977) declared himself clearly in favour of
sloping cores, arguing against what would be the symmetrical
solutions and their inconsistency in view of the asymmetrical
character of action by the water (although a vertical core
does not preclude a transverse cross section with asymmetrical
outer faces). 3 I, 4 between the risk of a failure in a
downstream zone with the reservoir full, and in an upstream
zone, due to a rapid drawdown, and this should not be covered
by only adopting different coefficients of safety. Failure of
393

the downstream zone ought to be virtually impossible, and in


fact the inclination of the core upstream, while keeping the
other factors equal, increases the safety of the downstream
zone of the dam.

/
N~ OF
DAMS

90 85 80 75 70 65 60 55 50 45

a (DEGREES)

Figure 18. Core inclinations on the dams of Table 1.

In the opinion of Mikuni (1980a and 1980b), vertical cores


with coarse, well graded soils are the best solution for good
behaviour against seismic action. In the case of clayey soils
with high water content, on the other hand, and for the same
kind of action, the vertical core solution is inadvisable,
particularly when it is very slender. The likelihood of arch
effects developing is greater. During the earthquake, the
unconsolidated part under the arched zone suffers a rapid
subsidence owing to vibration, and cracks that then appear may
give rise to concentrated seepage flows.
In the case of materials with a high water content, a
solution with a highly inclined core that is close to the
upstream face (Nantahala type) is recommended, but has no
longer been used owing to deformation of the core resulting
from the settlements of the principal rockfill shell. Such
deformations may cause undesirable stresses in the base of the
inclined core and give rise to tensile stresses during an
earthquake.
Cores that are slightly sloping (see Figs. 5 and 6) have been
increasingly used, mainly in very high dams, since they adopt
the merits of the two previous solutions while limiting their
drawbacks: deformation of the core is reduced by constructing
a very well compacted transition zone downstream, and the
394

upstream rockfill is used to transmit pressure to the core.


Subsequently, Pe?man (1983), Cooke (1984) and Sherard (1984,
1985 and 1986), 1n works devoted to rockfill dams, made no
statement as to what might be the best solution for the
position of the core.
The position of the core is thus a fact that may influence
structural behaviour and each case must be conveniently
analysed. The a priori option for one solution as the best one
is basically a matter of opinion.

L ___/ __- "'-


500 1000
kN I m 2

50m
i==I===Io1

Figure 19. vertical stresses at the end of construction. Homo-


geneous dam as regards deformability and shear strength.

3.3.4. Derformability of the Different Zones of the Dam. This


aspect is one of those that most decisively condition the
behaviour of the type of structure concerned. The different
deformability of the various zones of the body of the dam may
mean a transfer of stresses, and these may lead to undesirable
behaviours, i.e. they may bring the structure to ultimate or
serviceability limit states.
The ideal result would therefore be a structure which in its
operational phase showed a homogeneous distribution of
stresses. Fig. 19 gives the calculated distribution of ver-
tical stresses in such an ideal dam, at the end of construc-
tion. The different zones would correspond to materials with
different properties as regards permeability, filtration and
drainage capacity, but with equal properties from the point of
view of deformability and strength. It has been assumed that
the materials have a non-linear elastic behaviour and the
strength is sufficient for there to be stable in relation to
sliding along surfaces that concern the body of the dam. It
must be pointed out once more that in earth-rockfill dams,
rockfills, even of low-strength rocks, rather seldom raise
problems related with the shear strength.
395

and is therefore higher than the pressure due to the water of


the reservoir, which has to be not more than y wh (y w is the
water unit weight).
It is considered that the example pointed out by Sherard as
an extreme case is not as conclusive as the author seems to
indicate. A clay placed with a similar technique (wet compact-
ion) would show a better behaviour than a non-cohesive
material such as glacial till. In fact, the latter underwent
a rapid process of consolidation, and it is not surprising
that it should have developed an arch effect with transfer of
stresses. Is not this the behaviour which occurs in silos with
frictional materials?
As a final note on this important factor (the water content
of the core) it must be mentioned that an examination of the
dams referred to in Table 1 shows that, with the exception of
Emborca9ao and Chicoasen, the water content used in compaction
of the core is always higher than the optimum.
It must only be emphasised that in the Angostura Dam, the
lower half of the core was compacted with the water content
situated in the interval 0 to -2%, and in Chicoasen Dam, also
in the lower zone of the core and up to a quarter of the
height, the compaction was done with an optimum water content.
These are not orientations that can go against the general
tendency. First of all, they were applied only in part of the
core, and secondly, because in the zone of greater depth, the
state of higher stress to which such soil is subjected has a
favourable effect on its deformability (change of a dry soil
to a wet soil according to the terminology of the critical
states soil mechanics).

3.3.5. Filters. It is probably in this respect that there has


been recently made progress whose results have most influenced
and will still influence the conception of earth-rockfill dams
(Vaughan and Soares, 1982; Sherard, 1984; Sherard et al.,
1984a; Sherard et al., 1984b; Sherard and Dunnigan, 1985;
Kenney and Lau, 1985; Sherard, 1986).
Further knowledge of the various erosive processes in core
materials, and of the behaviour of filters (a subject whose
analysis in detail lies outside the aims of this work) had as
practical results not only the proposing of new criteria for
dimensioning filters but also, and mainly, a very important
change of attitude as regards structural conception.
In fact, since it has been recognised that the occurrence of
cracking inside the fills of dams is not uncommon (Casagrande,
1950), but principally since the verification of internal
erosion in the cores of dams attributed to hydraulic cracking,
a noteworthy effort has been made to try and propose steps for
preventing the occurrence of such phenomena. This effort has
been carried out above all in the fields of mathematical
modelling and constitutive laws, and in observation of the
real behaviour of dams, in some cases comparing the results
observed with those obtained from mathematical modelling.
396

500 1000 1500 200088


kNfm2 ~~

o 50 m
I I

Figure 20. Vertical stresses at the end of construction. Dam


with zones of different deformabillty and shear strength (for
the characterization of the materials of the zones 1, 2, 3 and
4, see Table 3 in section 4).

As may be seen from Fig. 19, there are no zones with concen-
tration of stresses, which are distributed gradually and
homogeneously throughout the dam body. The considerations
refer to a transverse cross section of the dam and assume a
plane strain situation. But they are also applicable to a
longitudinal cross section of the dam or to its tridimensional
behaviour.
But the reality is that not only permeability, but also
deformability and strength are not, in the natural compacted
particulate media used to construct dams, independent proper-
ties. This means that in order to ensure a sufficiently
impermeable core, it is impossible to make it resemble the
stabilizing shells (or the filters), from a deformability
point of view. As regards strength, on the other hand, it is
well known that, to put it in a very simplified way, it varies
in the same sense as deformabili ty , i. e. a material with
greater rigidity generally has a correspondingly greater
strength.
Fig. 20 shows the distribution of vertical stresses in the
above mentioned dam, this time in the hypothesis of the
properties of the materials in the varlOUS zones being
different as regards deformability and shear strength. The
respective values are presented in Table 3 in section 4. It
will be noted that there is a concentration of vertical
stresses in the rockfill shells that is due to a transfer
originating in the core (more deformable than the shells).
The filters or transition zones of granular materials are
also frequently zones of concentration of stresses. Recently
it has, in fact, been recommended that the compactness to be
required for them must not be very high (relative density of
397

not over about 70%), since this has little influence on the
respective filtering and draining characteristics but may
decrease rigidity in order to improve substantially structural
behaviour. strength diminishes, but in fact this is not very
important. One drawback that is sometimes pointed out is that,
in the hypothesis of seismic action, medianly compacted
filters may generate positive pore pressures owing to negative
dilatancy. Except in very special cases, this does not seem to
be a matter with practical repercussions.
Fig. 20 also shows the concentration of stresses in the
fil ters.
It is obvious that since the examples presented are based on
calculations, their value depends on the reliability of the
calculation model and on the values attributed to the princi-
pal parameters of that model. As an example, it may be
mentioned that in the calculation, no displacements are
allowed between the different zones of the fill, but if they
were, the results might be different, as is the case of the
Niikappu Dam in Japan (Mikuni, 198Gb). Fig. 21 shows the
distribution of the safety factors (relationship between
available strength and mobilized strength) after the filling
of the reservoir, in the hypothesis of there being no dis-
placements (Fig. 21a) and in the case of displacements being
permitted in the interfaces of the various zones (Fig. 21b).
It will be seen that when deformations are allowed along
the interfaces, there is a favourable change in the safety
factors. Deviatoric strains occur in the vicinity of the
separation between zones, and such deformations increase the
stability of the body of the dam.

,0

aI

bl

Figure 21. Distribution of the safety factors in Niikappu Dam.


a) without displacements along the interfaces between differ-
ent zones; b) with displacements in the interfaces (Mikuni,
1980b) .
398

It must also be mentioned that this behaviour (concentration


of stresses, and deviatoric strains at boundaries between
zones), have a counterpart in practice, and in fact there are
very many monitoring results that prove this.
Generally speaking, measurements of displacements (surface
or internal) are far more reliable than those of total
stresses. For this reason, together with a lower cost and less
complexity, observations of displacements are more frequent.
It is important, however, to emphasise that a relatively
homogeneous deformation process does not necessarily mean a
homogeneous distribution of stresses. Should there be zones of
different stiffness, a homogeneous deformability may in fact
mean that there is a concentration of stresses.
Knowledge of the installed stresses may permit an anticipa-
tion of the eventuality of the occurrence, at any stage in the
life of the dam, of tension cracks in the zones of cohesive
materials and of failures in any zone of the dam and for any
type of material.
The most significant aspects of the conceptions aiming at
minimizing the undesirable effects resulting from the inevita-
ble contrast of deformabilities are as follows:

a) Positioning of the core.

Adoption of a sloping core may prevent an important transfer


of stress from the core to the stabilizing shells. This
question has already been dealt with in 3.3.3 and on it there
is no consensus of opinion as regards indisputable and
generalized advantages. Preferably, the inclined core must be
supported downstream on a rockfill mass of low deformability
(which is obtained by using adequate compaction energy and
grain size). This is the case with Svartevann Dam (Fig. 6) and
Emborca9ao Dam (Fig. 8).

b) Increasing the stiffness of the core.

Clearly this aim is only reasonable if the core is more


deformable than the rockfill shells and this does not always
occur. Table 2 shows, for the case of the dams referred to in
Table I, a qualitative indication on the relationship between
the vertical deformations in the rockfill masses (b r ) and in
the core (be) for the construction stages, first filling and
after the first filling. As is habitual in some dams, the
first filling is superimposed on most of the constructional
phase. since the question of relative deformability between
the core and the stabilizing shells is being dealt with, it is
necessary to justify why Table 2 gives indications on dis-
placements. The fact of a core settling in the same way as the
rockfill masses during construction does not, however,
necessarily mean that they have the same deformability. If a
arch effect develops in the core, with the consequent transfer
of stress, this means that although with the same vertical
399

deformation, the deformability of the core in that direction


is higher than that of the shells. Accordingly, it is only by
measuring displacements and stresses that it is possible to
carry out an unquestionable quantification of the defor-
mabilities. That is to say, in some situations of Table 2
(when the displacements of the core are practically the same
as those of the rockfill zones) the relationship between
displacements may give no indication as to the relationship
between the deformabilities.
An analysis of Table 2 shows that during construction there
are cases in which the cores are more deformable that the
shells (Infiernillo, Scammonden, Talbingo and Monasavu) and,
wi th the exception of Talbingo, they correspond to cores
placed with high water content. On the other hand, the
Angostura and Emborcac;ao Dams and, to a far less striking
extent, Svartevann Dam, show an inverse behaviour (in Emborca-
C;ao Dam the core was placed with a water content that was less
than the Proctor optimum and the same occurred with the lower
half of Angostura Dam).
It is then necessary to consider Akosombo, Chicoasen, LG4,
Beliche and Caracol Dams, in which the cores will have a
deformability that is at least the same as the stabilizing
shells (in general the cores were compacted with a water
content higher than the optimum Proctor content).
This means that the most frequent situation will be that of
cores that are less stiff than the rockfill shells. It is,
however, interesting to note that during and after the first
filling there may be a modification of the relationship
between the displacements that was verified during construc-
tion. This alteration is due to an extremely complex process
in which an important part is played by phenomena of hydrody-
namic consolidation (core), collapse and creep (rockfill) and
the interaction between the core and the stabilizing shells.
In principle, a way of approximating the deformabili ties
would be to increase the stiffness of the core by raising the
specific energy of compaction. In reality, such a practice,
following the appearance of efficacious means of compaction,
is inadvisable because the material later has a more brittle
behaviour and this is undesirable for obvious reasons.
Table 1 does not refer to the specific compaction energy used
in the core, but in all the dams referred to in that table
that energy corresponds to the normal Proctor test, and the
reasons for this choice are basically those mentioned in the
preceding paragraph.
At other times the nature of the core material, namely the
grain size, for normal compaction leads to a fill of great
stiffness. This is the case with LG4 Dam (see Fig. 10), where
the core, consisting of a glacial till, underwent extremely
limited settlement (19 cm), although the height of the dam is
125 m.
400

Table 2. Relationship between vertical displacements


(settlements) in the rockfill shells (Sr) and in the
core (Sc) in the dams of Table 1.

Relattonshlp between settlement


In rockfi[ls and cores (6r/05c)
DAM Remarks
construct IOn 1npoundIng

,
construct Ion
INFIERNILlO
" " >1 <1 after 10 years

AKOSOMBO ~1 >1

3 SCAMMONDEN ", " (') =1 4 years for reservoIr


fIll mg
TAlBINGO ", " "
ANGOSTURA , >, >,
6 SVARTEVANN >, >1 Main part of mpoundm9
dJrm9 constructIon
CHICOASEN ~, ~, ~,

8 EMBORCAI;AO , >1a) <lb) a} upstream shell


b) downstream she L I

MONASAW ", '<,

'0 L04 ~, ~,

BEL ICHE ~, >, >1 a) a) arch effect


"
,2 CARACOL ~,

Owing to fear of the high pore water pressures that occur


during the construction stage (a fear that is only j ustif iable
in cores that have a marked inclination), cores were even
constructed and compacted on the dry side (with a water
content lower than the optimum). This technique, however, is
open to criticism because the material becomes brittle and
less capable of adapting to differential settlements.
Another procedure consists of compaction on the dry side in
the deepest zones of the core and a rather wet compaction at
the upper elevations (the case of Angostura Dam, Fig. 5). A
stiffer core is thus obtained in its lower zone, with attenua-
tion of the phenomenon of transfer of stresses to the rock-
fills. But since the upper zone is very prone to cracking,
compaction is on the wet side in order for that zone of the
core to have a more ductile behaviour, i.e. the core transfers
more stresses but its material is more likely to be deformed
without cracking.

c) Reducing the stiffness of the shells.


Another possible way of homogenizing stresses through
uniformization of the deformabilities in the various zones of
the structure is to reduce the stiffness of the rockfill
401

masses. In principle, the negative consequences from the point


of view of strength are not significant, and the same holds
good as regards permeability. Even though they do not have
that preoccupation, this is the kind of solution used in
earth-rockfill dams that were constructed before the advent of
vibratory compaction techniques, as is the case with the
already referred Cherry Valley Dam shown in Fig. 11. Of note
is the great thickness of the core, thus showing a slenderness
of 2, which is therefore much higher than the maximum of 0.63
indicated in Table 1. Later came projects with zoning of the
rockfills such as Infiernillo Dam (1963), in which the core is
surrounded by compacted rockfill and there is an outer cover
of dumped rockfill (Fig. 1) and Cougar Dam (Fig. 22), in the
United states, completed in 1964 (Pope, 1967). Observation of
the typical cross sections shows that in the latter case the
core is inclined and, unlike Infiernillo Dam, only the
rockfill zone adjacent to the core and situated downstream is
compacted with vibration.

511 50
-~--

o50rn
~

-CLAY CORE SPALlS

- GRAVEL TRANSITION 5' SOUND ROCKFILL (COMPACTED WITH De)

~ ,
-LOW QUALITY ROCKFILL {COMPACTED WITH D8) t 6 - SOUND ROCKFIlL (COHPACTED BY VIBRATION)

Figure 22. Cross section of Cougar Dam.

There are, however, zonings in which the rockfills sur-


rounding the core are more deformable (mainly owing to the
quality of the rock) than those of the outer zone. This is the
case with Beliche Dam (Fig. 11), where the aim was to make use
of an available weathered material and, if there were a
transfer of stress, to ensure that such stresses would take
place from the stabilizing shells to the core, thus diminish-
ing the probability of an occurrence of hydraulic cracking.
402

Also from a seismic point of view this conception may be


advantageous in relation to the first mentioned solution.
It is to be noted how sometimes the desired objective
(reduction of the stiffness of the rockfills) can only be
attained later (after construction) and the cases of In-
fiernillo and Beliche illustrate this precisely in Table 2.
The fact is that it is only in the operational stage that
these dams reveal greater displacements in the rockfill
shells. This is due to settlements by collapse of the upstream
rockfill (due to submersion on account of the rise in the
water level of the reservoir) or of the downstream rockfill
(due to rainfall, principally if the type of rockfill justi-
fies the use of water during its placement and that water has
not been added), and also to settlement resulting from creep.
Creep, of course, depends largely on the degree of compactness
(or on the stiffness reached during compaction), and this
explains the greater deformations which, in principle, appear
in dumped rockfills (or rockfills compacted in very thin
layers - 0.6 and 1.2 m - but with the use of a very low
specific compaction energy).
As disadvantages of this type of structure, i.e. with
relatively deformable stabilizing zones, mention must be made
of the high probability of the occurrence of longitudinal
cracks in the zone of the crest. This behaviour is well
exemplified in all the dams referred to (Infiernillo, Cougar
and Beliche), which have revealed that type of surface
cracking. Cherry Valley is an example of an earth-rockfill dam
that also has highly deformable rockfill masses, but it date
from the period before vibratory compaction, and this dam also
showed longitudinal cracking on the crest.
Fig. 23 presents a scheme of the longitudinal cracking in
Cougar Dam that illustrates how the cracks are associated with
the interfaces which establish the zoning.

120 m LONG CRACK,


MAX. 015 m VERTICAL
OFFSET
90 m LONG CRACK
MAX. 0.30 m VERTICAL OFFS.~E~T::;:--=--~WrJf=r--i":: - - DOWNSTREAM
SHOULDER

ROCK

Figure 23. Scheme of the longitudinal cracks in Cougar Dam


( Pope, 1967).
403

Mention must also be made of Emborca9ao Dam since, as may be


seen from Table 2, there is an inversion downstream, during
filling, of the relationship between the displacements of the
core and of the rockfill. But since upstream there was
maintained the initial relationship (rockfills more deformable
than the core) longitudinal cracks appeared on the crest,
associated with differential settlements: the upstream
rockfill settled 20 cm in relation to the core and the core
settled 2 cm in relation to the downstream rockfill. This
behaviour is not very different from that one observed in
Cougar Dam.
In Fig. 24 are to be seen the longitudinal cracks that
occurred in the upstream side of the crest of Beliche Dam
(similar cracks were observed in the downstream side) as the
result of differential displacements whose horizontal compo-
nent brought about an extension of elongation of 1.5%. The
displacements, all of them in a downstream direction, took

Figure 24. Longitudinal crack in Beliche Dam (upstream side of


the crest).

place between June 1986 and May 1989, when the water level in
the reservoir rose from half height to maximum height (LNEC,
1989) .
An interesting aspect from the point of view of conception
is the use of a clay blanket prolonged downstream, at the base
of the inclined core (v. de Mello, 1983) and exemplified in
404

Fig. 8. It is the particular case of the vertical and subver-


tical cores with widening at the base (see Infiernillo Dam, in
Fig. 1, for example). The purpose of this solution is to
reduce the gradients of seepage in the foundation and intro-
duce a vertical deformation in the rockfill masses (stiffer
than the core) that will make the settlements similar in the
core and in the rockfills surrounding it, and thus minimizing
the transfer of stress. In practice this solution has not been
much used, and when it was used, monitoring of behaviour has
not clearly backed up the assumptions which, as regards the
contrasting deformabilities, were the basis for its use.
Lastly, it must be pointed out that, contrary to the traJ"ls-
verse cracks, the longitudinal cracks on the crest do not
affect the safety of the work. They do not lead to ultimate
limit states but tend to bring about serviceability limit
states (road on the crest unutilized, aesthetic inconveni-
ences, operationality of the dam affected by repairs to the
crest, etc.).

d) Reducing the stiffness of the filters or transition zones.

This matter has already been referred to at the beginning of


this section, and nowadays there is little argument about the
orientation whereby the filters must be medianly compacted in
order to prevent a concentration of stresses in those compo-
nents.

e) Use of a high water content in the core.

This procedure is intended to counter one of the most perni-


cious effects due to a core being more deformable that the
surrounding rock masses - hydraulic cracking. It is often
advocated (Penman, 1983) and occurs naturally owing to a
constructional technique that has been used in British dams
since the end of the last century (use of puddle clay), a
technique that is very compatible with the high natural water
content in the clays used in the dams.
The scheme in Fig. 25 indicates that the total stress (a) of
the clay of the core on any plane which contains point P must
be higher than the pore water pressure (u) at that same point.
In fact, unless that is so, a crack forms along a plane normal
to the minimum principal stress (the effective stresses on
that plane are null).
Since the undrained shear strength of the clay of the core
(c ) enables part of the stress due to the self weight of the
co~e to be transferred to the surrounding rockfill masses, it
happens that the vertical stress (ov) may be considerably less
than the overburden stress. Obviously the horizontal stresses
(oh) have even lower values. A very simple formulation will
give, provided that we assume a situation of integral mobili-
zation of the undrained strength of the core material, the
40S

following:

so that the total horizontal stress will be

where y is the total unit weight of the soil of core, b is the


width of the core and Ko is the coefficient of soil pressure
at rest.

Figure 25. Total and pore water pressures in the core of an


earth-rockfill dam.

In general, the higher the strength of the material, the


lower is the value of Ko ' ~o that in cores of clay with a high
c u ' there may be total horlzontal stresses of very low values.
If they act according to the direction of the axis of the dam,
there is a high probability of a hydraulic cracking along a
transverse vertical plane, i. e. with the most unfavorable
orientation from the point of view of the consequences of the
phenomenon being analysed. That is to say, there may be
established a preferential hydraulic connection between the
upstream and downstream shells, probably leading to a process
of internal erosion of the core.
If the water content used in compaction is much higher than
the optimum, there is no possibility of using normal compact-
ing equipment and men's feet are then used to try and homoge-
nize the material. Swedish and Norwegian experience is also
known in the compacting of their dam cores, generally con-
structed of glacial alluvia (till), with water content so high
that they can only be compacted with crawler-type tractors.
This technique is known as "wet compaction" (Sherard et al.,
1963). The compactness thus obtained is naturally far inferior
406

to that corresponding to the maximum obtained in the normal


Proctor test, but the permeability is nearly the same as that
which would be attained if the fill had been constructed with
the use of heavy compacting equipment and a water content
equal to the optimum of the Proctor test. One of the best
known examples is that of the Swedish dam of Messaure, built
in 1962, with a height of 101 meters and a vertical central
core consisting of glacial till.
As might be expected, the cores of these dams later undergo
considerable settlements owing to consolidation. Such settle-
ments may be very quick. In cases of plastic soils, low
ground-pressure tracked D6 bulldozers were used, so that they
would not sink into the ground. This is the recent case of
Monasavu Dam (see Fig 2.9), completed in 1982, where a wide-
track D6 tractor "compacted" layers of 0.1 m of a soil which
showed an index of plasticity of about 50%. The water content
was 70% and the clay showed a mean undrained shear strength of
17 kN/m2. During construction, the core had a maximum settle-
ment of about 1.6 m (half height) against a maximum settlement
of 0.3 m in the downstream rockfill. According to Knight et
al. (1985), the overall indication is that, not withstanding
the possible hang-up of the much stiffer filter, the soft clay
has squeezed down satisfactorily over about at least a 12 m
horizontal width.
As regards high pore water pressures, only in special cases,
as has been said, can they affect stability. But they can
prevent hydraulic cracking during the first filling (a
critical phase as regards this phenomenon). In fact, for an ru
(defined as the ratio u/o v ) at the end of construction or just
at beginning of the filling of the reservoir, of about 0.5,
for example, there is a low probability of hydraulic cracking
in the final phase of the filling because the pore water
pressures will tend to go down to the values corresponding to
the seepage flow that will be installed in the core owing to
pressure from the water in the reservoir (in Talbingo Dam the
value of ru measured at the base of the core at the end of
construction was 0.95 and ten years later 0.65; In Monasavu
Dam an r~ of 1.14 was even measured during construction, but
before fliling it had fallen to 0.5).
Since

and, according Fig. 25, 0v ~ (h + h') y and y (total unit


weight of the core material) is practically equal to 2 Yw' the
pore pressure in the fill is thus

u 0.5 (h + h') y (h + h') Yw


407

For a soil that can induce high pore water pressures during
construction, it is normal for Ko to assume very high values
(close to unity), so that 0h would not differ greatly from 0v.
In the quantification of r u' 0v is taken to be practically
equal to the overburden pressure, which means that there is no
transfer of stresses to the rockfill masses. This hypothesis
may not occur for cores of soils with moderate to high shear
strength, so that the consequent diminution of 0v implies also
a diminution of ru. This means that the reductions in pore
pressures so often observed during construction may be mainly
due to a transfer of stresses to the stabilizing shells and
not to a hydrodynamic consolidation. The shear strength of the
core material must therefore not be very high in order to
ensure a very deformable core, likely to develop at the
interfaces with the foundation (including abutments), with the
stabilizing shells and on any plane that intercepts it, total
stresses always higher than the pressure in the water due to
the reservoir. In this situation, the horizontal stresses
applied by the core on the upstream (and downstream) stabiliz-
ing shells must be higher than those of the water in the
reservoir on the core. Accordingly, the pressure of the core
on the downstream shell does not undergo any appreciable
alteration and, during filling of the reservoir, there are no
displacements of the core downstream.
According to Penman and Charles (1973), this was the
behaviour observed at Scamrnonden Dam.
From what has been said, it may be concluded that a good
structural solution is that which results from the combination
of a very deformable core, low shear strength and developing
high pore pressures during construction (which ensures
watertightness of the body of the dam and prevents hydraulic
cracking problems) with very steep compacted rockfill masses
(which will ensure stability as regards failure by sliding in
the body of the dam). It will be a safe and economical
solution as regards the overall cost of the materials to be
placed (Fig. 26).
This solution has, however, certain drawbacks, one of which
is due to the effect of time (Sherard, 1985). During the life
of the structure, the initial water content may be modified.
If the operation system includes situations of filling after
fairly prolonged periods with the reservoir at a very low
storage level, during that filling hydraulic cracking may
occur.
Sherard (op. cit.) even points the case of Messaure Dam,
already referred to, in which, after the first filling, there
were considerable concentrated seepage through the core
causing its erosion. It was concluded that the unsatisfactory
behaviour was due to arch effects that were generated in the
core, that is to say, even a compacted glacial alluvia (till)
with high water content may transfer stresses.
40X

1.5
1

Figure 26. Cross section of an earth-rockfill dam with a core


compacted with high water content dnd sound well compacted
rockfill in the shells.
The procedure consisted of calculating stresses in the body
of the dam for different geometries and different design
situations and, taking as background the mechanical and
hydraulic properties of the fills and the knowledge obtained
from monitoring the dams, deciding on the structural concep-
tion that would be likely to meet all the other requirements
of design behaviour and prevent the installation of limit
states (either ultimate or serviceability) directly connected
with fissuration and hydraulic cracking.
In this strategy the filters acted as a second line of
defence, i.e., if the dam, even though designed according to
the procedures described above, later developed internal
erosion, the filters would act as a controller of that erosion
and keep the dam in a safe condition.
According to Sherard (1984), nowadays it is possible to
introduce an important alteration in this strategy. The
filters, or more specifically the filter adjacent to the
downstream slope of the core (known as critical filter),
constitutes the most important and efficient line of defence.
Provided that it is properly dimensioned and constructed, it
controls internal erosion of the core caused by cracking,
namely erosion originating in hydraulic cracking, thus
ensuring the safety and serviceability of the dam.
This means that emphasis as regards safety is not placed on
analysis of the stress-strain behaviour and on the many
measures for reducing differential settlements (or concentra-
tion of stresses), but rather on dimensioning of the critical
filter.
It may be asked why, in the fifteen odd dams which since
1955 have registered concentrated seepage through the core and
intense internal erosion, the filters have not controlled such
erosive processes. Such behaviour has been due to a dimension-
ing which, in the light of the most recent research in this
409

field, has proved deficient. In fact, all the cores of those


works consisted of broadly graded soils, represented approxi-
mately by a straight line in the habitual semi-logarithmic
representation (normally saprolitic soils or soils of glacial
or colluvial origin) and only recently have auto-filtration
studies in general (Sherard and Dunnigan, 1985), and in
particular studies on non-cohesive broadly gLdded soils
(Lafleur et al., 1989) and on compacted clayey soils (Maranha
das Neves, 1989) allowed a clarification of this phenomenon
with such far-reaching repercussions on the behaviour of
filters.
The growing importance of the role played by critical
filters concerning dam safety has been followed by the
progressi ve devaluation of the similar contribution of the
filters placed upstream of the core. Mainly due to the fact
that only very low gradients can occur on these filters, they
can be built with a single band of a relatively coarse
material (a grizzled hard quarried rock, for instance) even
for the transition between a core of fine clay to the upstream
rockfill shell. This is really a very liberal option.
Finally the use on the top of the upstream face of the core
of fine uniform sand filters acting as crack fillers is
debated nowadays. According Sherard and Dunnigan (op. cit.)
this solution will be less used in the future and probably
will be abandoned as a design practice.
3.3.6. Shape of the valley. In the foregoing considerations,
the predominant analysis have been based on bidimensional
situations, corresponding to plain strain states, for example,
the case of structures with constant cross section and
infinite longitudinal length.
In fact, however, real conditions may differ considerably
from those assumed and, principally in the case of narrow
valleys, the tridimensional effects may be import.
Table 1 gives data on the relationship between the longitu-
dinal extent of the dam (L) and its height (H), for the twelve
dams analysed. It will be noted that the L/H ratio ranges from
1.2 in Chicoasen Dam (261 m high) to approximately 10 in
Beliche (54 m high) and Emborca9ao (158 m high), which denotes
a wide range of variation in the L/H value.
According to Lefebvre et al., (1973), L/H ratios lower than
6 lead to significant tridimensional effects. One of the main
consequences of a narrow valley is the occurrence of longitu-
dinal deformations towards the zones where the height of the
dam is greater, decisively influencing the stresses normal to
the transverse cross sections. If the valley is not only
narrow but also asymmetrical, deformations may occur (during
construction or in the operational stage) in the approximate
direction between the sides of the valley, but with one
predominant sense. The stresses associated with these deforma-
tions are hard to foresee but they may often influence the
410

structural behaviour. Another important aspect is the develop-


ment of arch effects, with transfer of stresses from the body
of the dam to the abutments.
An exemplary case, also owing to the height of the work, is
Chicoasen Dam. Fig. 27 gives a longitudinal cross section of
the dam, showing the zones of the core adj acent to the
abutments which are compacted with a water content of about
+3% in relation to the optimum content. These arrangements,
together with zones of the core that are adjacent to both the
upstream and downstream shells and placed with a water content
of +3% to +4% (see Fig. 7), were intended to attenuate as far
as possible the effects of interaction. Even so, measurements
indicated a marked interaction between the core and abutments
and stabilizing shells. The result was an arch effect. This
conclusion was reached when the pressure cells placed at the
core foundation level registered vertical stresses that were
far lower than the geostatic stresses. This end-of-construc-
tion condition might well lead to hydraulic cracking, but
monitoring of the dam has not detected that phenomenon or any
anomalous behaviour of the structure. During filling, it was
found that the very wet material of the contact zones under-
went shear rupture, thus helping to reduce the core-abutment
interaction generated during construction.

LB RB

w
150.00
o 50 100 m
-=-; I I I

1 T -Clay material (at optimum)

1 C -Clay material (0.8 OJ. dry of optimum)


lW-Clay material (2 to 3"10 wet of optimum)
L8-Left bank
RB-Right bank

Figure 27. Longitudinal cross section of Chicoasen Dam (Gon-


zalez Valencia and Aguirre Soria, 1985).
411

Fig. 28 shows the asymmetrical valley in which Svartevann


Dam was constructed, and the zones where there is a tendency
for development of tensile stresses related with the shape of
the valley. The same figure represents graphically the
displacements measured on the strain meters six years after
completion of the work. The length of the strain meters varies
from 9 to 12 m, so that the maximum strain due to tension is
about 0.1%. Fig. 29 refers to the surface displacements
tangent to the axis of the dam, on the same date. It will be
noticed that these converge to the axis of the valley, and the
general configuration is of approximate symmetry in relation
to that axis, although it represents the reduced degree of
asymmetry of the valley.

oj

~wm
.~Z2/'z;Z;Z;VZl

..z>2J22WaU2J

~
mm z
I-J ~

100 a..
~ 8~6.50' 877.3~
<J) ,

0 ./
- -=~.J,;~r500
I. I_COMPRESSIVE STRAINS

1-1- TENSILE STRAINS

a 100 m
~

Figure 28. Tensile stresses in Svartevann Dam (Di Biagio et


al., 1982) a) strains measured b) location of the strain
meters in the longitudinal cross section of the dam.

In some cases, the soil of the core has been placed with
high water content in those zones where tensile stresses are
expected develop, as is the case with Sklope Dam, built in
1967 in Yugoslavia, reported by Sherard (1973) and repre-
sented in Fig. 30. A reinforcement consisting of a geomesh of
polymer material can be used, as is the case with Canales Dam,
built in 1986 in Spain (Guillen Bravo et al., 1988).
About 20 years ago some interesting investigations were
carried out in order to establish a correlation between the
type of soil and the potential of dams for transverse cracking
412

(Sherard, 1953; Leonards and Narain, 1963; Biarez et al.,


1970). Those studies did not continue because it is very hard
to arrive at generalized conclusions of practical utility, and
in fact it is possible only to infer some tendencies.

Figure 29. Superficial horizontal displacements (in meters)


tangent to the dam axis in Svartevann Dam (Di Biagio et al.,
1982) .

Zones of clay core placed


at high water content to increase
ability to stretch without cracki ng
.~---="'<=

o 20m
~

Figure 30. Compaction on the wet side in zones of Sklope Dam


where tensile stresses were expected to develop (Sherard,
1973).

Lastly, reference must be made to cases in which the


abutments, even though locally, show high variations in
inclination. The consequent differential settlements may be
the cause of cracking in the core, especially during the
construction stage. This is the case with Round Butte Dam,
built in 1964 in the United States, with a height of 132 m. As
can be seen in Fig. 31, a transverse crack was located in the
vertical of the prominent irregularity of the rocky abutment.
In other cases the irregularity configuration implies its
filling of with concrete (Fig. 30).
To sum up, all the above concerns repercussions of the
413

geometry of the valley on the development of states of stress


leading to the appearance of fissures or hydraulic cracking.
This means that when tensile stresses occur in the cohesive
soils of a core, there is a high probability of cracking
taking place and if, with the filling of the reservoir, the
effective stresses in the core (or in its contacts with the
foundation rock or with the hydraulic structures) are an-
nulled, hydraulic cracking will occur. The tensile strength of
cohesive soils and the mechanisms which characterize the
phenomena of hydraulic cracking are controversial matters in
soil mechanics, but there is no doubt as regards their
eventual consequences, i.e. the occurrence of internal erosion
in the core of a dam.

Crest. 596

Transverse crack

o 15 m
I I

Figure 31. Transverse crack in Round Butte Dam due to high


variations of inclination of the abutments (casagrande, 1965).

Also in this case the critical filter is an important factor


that makes it possible nowadays to affirm that narrow valleys
do not constitute an impediment to the construction of an
earth-rockfill dam. Moreover, in valleys with a configuration
in which, until recently, the only solution would have been a
concrete arch dam, fill dams can now be built.
Apart from the important part played by the critical filter
in dam safety, it is current and accepted practice today,
whenever the geometry of the valley makes it advisable, to
built a strip of fill of about 3 m adjacent to the founda-
tions, the abutments and the hydraulic structures, with a
placement water content that is 3% above the optimum.
Another important aspect are the geometric characteristics
of the surfaces of the hydraulic structures in contact with
the core. The normal practice of designing walls protruding
414

out into the surrounding dam core (to increase the seepage
path) must be given up. The surfaces must be smooth and sloped
slightly off vertical (slopes about IH:IOV) so that the earth
can be compacted directly against the concrete by a rubber-
tired vehicle running parallel with the concrete face, which
needs not to be sandblasted, chipped, painted or treated in
any way.
By similar reasons seepage collars of concrete on conduits
passing under the core (also a standard practice) are of no
use.
Nevertheless a special care is recommended on the above
mentioned situations: all potential leakage water travelling
along the concrete-core interface must exit through a filter.
Thus any internal erosion will be controlled.
The reasons for all these changes is that observed behaviour
shown that the inconvenients of bad compaction inherent to the
complicated geometry of the interfaces are higher than the
disadvantages of having lower seepage paths.

3.3.7. configuration in Plan. As regards the configuration in


plan, we shall deal only with the case of the geometrical
shape of the longitudinal axis of the dam, owing to its
eventual influence on the structural comportment of the work.
It may be straight (commonest solution) or curved with
concavity downstream (see Fig. 32, showing the plan of Beliche
Dam) .

Figure 32. Plan of Beliche Dam.


415

It is frequently argued that under the action of the


reservoir water, a curved axis would, owing to arch effect,
tend to cause an increase in the stresses acting in the
direction of the longitudinal axis, thus reducing the poten-
tial for cracking owing to tensile stress, mainly transverse
cracking.
In 1963 Sherard et al. were of the opinion that the advan-
tages of a curved axis are not significant, but in view of the
limited extra cost of this solution those authors envisaged an
increasingly generalized use of curved axes, but in fact their
forecast has not been borne out.
Twenty years after Sherard (1984) stated that he will be not
in favour of arching the dams (upstream convexity) but he will
be not against if somebody takes this decision.
Penman (1985) questioned the effectiveness of arching
embankment dams in plan to raise longitudinal stresses as a
result of the downstream crest movements due to the reservoir
water pressures. He argued that there is no measurements of an
increase of the referred total stresses conjugated with a
downstream movement.
According to Sherard (1985) a final (after impounding)
downstream movement of some centimeters implies that the crest
length must decrease (compress) by a small amount.
The main question is the value of the compressive longitu-
dinal stresses associated to these displacements. They are
probably very low so difficult to measure even with reliable
equipments. And if they were measured there would be no way to
know what the stresses would have been in the dam if it had
been constructed with a straight axis (Sherard, op. cit.).

3.4. FINAL REMARKS

What has been stated above gives some account of the develop-
ments in the conception of earth-rockfill dams, and thus tries
to make current recommendations on the conception of this kind
of work clearer and better founded.
Given below are the most significant aspects of conception,
related with geometry, deformability and seepage, with the aim
of good static structural behaviour of earth-rockfill dams.

a) critical filters contribute decisively towards safety as


regards the effects of cracking and hydraulic fracture.
b) As regards the static behaviour of the structure and in
terms of transverse cross section, there is no ideal and
exclusive positioning for the core.
c) The water content for compacting the core materials must
be close to the optimum or slightly higher that the
optimum, and a specific compaction energy corresponding to
the normal Proctor test should be used.
416

d) If the core materials have a very high natural water


content, it is possible, and admissible, to use the wet
compaction technique.
e) It is not necessary to insist on clayey plastic materials
for the core, but only that the materials shall be
permeable enough for the purpose concerned.
f) In principle, there is no limitation as regards the type
of rock to be used in the rockfills of the stabilizing
shells (except in the case of materials of high weather-
ability, when the type of rock will have to be considered
in the light of the particular conditions concerned). In
fact, the use of low-strength rock materials may be
considered current practice. Weak rocks (taken to be
transition material between rock and soil) will normally
give raise to fill to which can be applied the principles
of soil mechanics, i.e. they cannot be regarded as
rockfills.
g) By the foundation (including abutments), the core material
must be compacted with a water content that is definitely
higher than the optimum content (~ +3%). This principle is
also applicable to the contacts of the core with the
hydraulic structures. In special cases it may be used in
the contact of the core with other zones of the fill.
h) The filters adjacent to the core must be moderately
compacted in order not to have a stiffness that is in
great contrast with the surrounding fills, and thus
leading to high concentrations of stresses.
i) The limitations to slenderness of the core are related
with the eventual appearance in the core of significant
arch effects.
j) In principle, the geometry of the valley cannot give rise
to any impediment to built earth-rockfill dams.
k) Seepage collars on conduits and at core-concrete struc-
tures interface are no longer used.
1) In principle, curved longitudinal axes do not offer any
appreciable advantages over the rectilinear axis.

4. Modelling the Structural Behaviour

What has been explained in section 3 above indicates that


al though there are nowadays some principles and rules with
which the structural conception of an earth dam must comply,
there is still a long way to go before an essentially scien-
tific approach to the question is achieved. The quantity of
parameters concerned is such that the empirical component and
experience continue to play a fundamental part in the dimen-
sioning of this type of structure.
As an example, and taking as basis the results of monitoring
of the dams referred to in Table 2, Fig. 33 shows the maximum
417

vertical strains occurring in the cores during construction,


according to the height of the dam. It will be noted that the
dispersion of the values is very great (0.3 to 4.4%), and
there is no correlation with the height. The most that can be
said is that the greatest vertical deformations developed in
those cores that were compacted with very high water content
("wet compaction"), which in fact might be expected.

Table 3. Physical and mechanical characteristics of


Selimo Dam materials.


'.
y c H I PER B 0 L I C PARAMETERS
,
Y sat
MATER] Al
k~/m3 kN/m 3 kN/m Z degrees k n Rf G f d kur

,. CLAY 21.7 77 20.1 180 0.65 0.90 0.40 0.10 2.1 1080

2- SAND AND 17.9 , 47.0 1500 0.15 0.87 0.35 0.41 10.6 9000 12.5
GRAVEL

18.2 0 5'.0 525 0.1 0.75 0.18 0.18 8 3150 13.0


3 SCH I ST AND
GREYI,./ACKE
(I,./EATHERED)
20.3 , 37.0 135 0.36 0.75 0.18 0.18 8 810 0.0085 4.0

20.5 0 48.5 340 0.6 0.78 0.24 0.15 12 2040 7.0


4-SCH[ST AND
GREY\.IACKE
(SaUND)
23.1 a 45.0 150 0.6 0.68 0.24 0.15 2 900 0.0026 5.0

Reference must also be made to the recent works by Clements


(1984) and Dascal (1987), which analyse the settlements after
construction of a large number of dams. They indicate some
behaviour trends for earth-rockfill dams, but even so these
are limited to a certain type of dam, and they raise doubts
about the practical value of some correlations previously
presented with general applicability, by certain authors. This
attitude is expressed by Dascal (op. cit.) in a clearer form
and it has been corroborated by other authors (McCleskey, 1988
and Nonveiller, 1988).
This enormous complexity and diversity of structural
behaviour of earth-rockfill dams signifies precisely that the
theoretical approach has its place in structural dimensioning.
It thus plays a role which is not unique but which can
nowadays be regarded as irreplaceable.
The scientific approach means predicting dam behaviour by
mathematical reasoning on the basis of test results and a few
fundamental relationships.
In the present case, use will be made of the hyperbolic
model, a description of which can be found elsewhere (Naylor
et al., 1986, for example) and it will be applied to a
fictitious dam with a height of 120 m to be designated Selimo.
Calculation of the stresses and deformations was done by
means of a finite element programme developed at LNEC (Mateus
da Silva, 1990), using isoparametric elements of eight nodes.
418

280

H 7
D
1m)

240

_SOME CORE CHARACTERISTICS_


200
20; _ SLOPING CORE
_COMPACTION ON DRY SIDE
o _ COMPACTION ON WET SIDE
4; @) _ "WET COMPACTION"
160 8.i 0
1 5 fj) _HIGHLY PLASTIC
0 D
~ _ TI LL
6; D _ ONLY UPPER ZONE IN WET SIDE
10 12
0 ~
120 ~
2
>

9
80 3;
@)
11
0

40

EV (%)

Figure 33. Maximum vertical strain (Ev) occurring during


construction in the dams of Table 1 according to dam height
(H)

On the basis of a parametric study, comments will be made on


the structural behaviour deduced from this modelling, taking
into account experience of the real behaviour of this type of
structure. From a geometrical point of view, two dams will be
considered: Belimo I and Belimo II. The first has a vertical
core and the second a sloping core.
The materials are the same as those used in Beliche Dam
(Fig. 11) (Veiga Pinto, 1983). In fact, apart from the
alluvial foundation and the upstream cofferdam, the geometry
is much the same as that of Beliche Dam. The values of the
physical and mechanical characteristics (hyperbolic parame-
ters) of those materials appear in Table 3. The parameters
provided make it possible to take into account curved failure
envelopes (a) and to describe the behaviour when unloaded
(Kur ) ' Also taken into consideration in the model is the
phenomenon of collapse during filling of the reservoir
(Maranha das Neves and Veiga Pinto, 1988).
419

4.1. DAM WITH A VERTICAL CORE

Figs. 34 and 35 show the cross section and the vertical


stresses at the end of construction of Belimo I Dam, respec-
tively. Some arching can be observed in the core, with stress
concentrations in the filters at very low levels.

120.00
~

o 50 m
I (

(2) - CLAY 0) - SCHIST AND GREYWAKE (SOUND)

CD - SAND AND GRAVEL CD -


ROCK (FOUNDATION (

0- SCHIST AND GREYWAKE (WEATHERED)

Figure 34. Cross section of BeL:mo T Da.m.

500 1000 1500 2000 ~ 3


k N 1m2 "'0
"'''>
50 m
1=====11

Figure 35. vertical stresses at the end of construction of


Belimo I Dam.
420

After filling of the reservoir, total vertical stresses are


obviously higher in the upstream shell (Fig. 36) and remain
almost unchanged in the downstream shell. In the upper part of
the core the arch effect increases and the stresses grow in
the filters, mainly in the upstream filter.

o 50 m
I I

Figure 36. vertical stresses in Belimo I Dam after filling of


the reservoir.

One interesting analysis consists of changing the materials


between the two zones in the shells: the weathered rockfill
goes outside and the sound rockfill is placed inside. The
results can be observed in Fig. 37 and it is evident that when
compared with those of Fig. 36 stress concentrations in the
filters disappear, as a result of the surrounding sound
rockfill. On the other hand, vertical stresses in this zone of
the shell are higher when its deformability is lower.

1000

o 50 m
I I

Figure 37. Vertical stresses in Belimo I Dam after filling of


the reservoir but putting t.he sound rockf i lIon the inner
shell.
421

stress concentrations in the filters underline the extreme


and well known influence of deformability of the filter
material on stress distribution. In order to bring out this
fact, vertical stresses after reservoir filling were calcu-
lated assuming a more deformable filter material. The adopted
values are K=700, Kur =4200, =35 and 1l=8 (cf. the values for
the same parameters given in Table 3) and the results are
presented in Fig. 38. Comparison with Fig. 36 shows that
vertical stresses in the core are higher for more deformable
filters and stress concentrations in the filters (mainly in
the upstream filter) practically vanish. A relevant aspect is
that core vertical stresses with the deformable filters are
very similar, at least in the lower part, to those obtained
with normal filters but with the sound rockfill in the inner
zones of the shells (Fig. 37). This highlights the importance
of the relative deformability.

1000

2000 2500 3000 2000 2500

o 50 m
t=====i

Figure 38. vertical stresses in Belima I Dam after filling of


the reservoir but with a less stiff and resistant filter
material.

Another very stimulating analysis is comparison of the


horizontal stress (oh) in the core after construction and
after filling of the reservoir. This was done for the two
structures: one with the weathered rockfill in the inside
zones of the shells and the other with this material in the
outside zone.
Fig. 39a and 39c show the 0h values in the cores and they
are a little higher where the weathered rockfill (Figs. 39b
and 39d) it is outside. But after the filling of the reservoir
it is the dam that has have the weathered rockfill in the
inner zone (Fig. 39b) that shows higher stresses, and this
means lower probability of hydraulic fracture.
Using a different constitutive law, Naylor (1990) obtained
results for Beliche dam that show stress transfer from the
shells to the core that was attributed to placement of the
422

weathered rockfill near the core. (Belimo dam uses practically


the same solution as in Beliche Dam). These conclusions are
well supported by the results of the dam monitoring excluding
part of the foundation. The collapse due to submersion of the
rockfill as well as the reservoir water pressure explain why
the stresses are higher in the upstream zone of the core (see
Figs. 39.b and 39.d). In the case of Beliche dam, settlement
of the downstream shell was also due to collapse caused by the
rains. As a consequence the stresses in both the upstream and
downstream core zones are more or less similar.

a) b)

750
1250
1000

c) d)

a 50 m
I I

Figure 39. horizontal stresses in the core of Belimo I Dam


a) er~J. of constructiofl, weathered rockfill on the inner
shell;
b) ar'Cer filling of 'che reservoir, weathered rockfill
on the inner shell;
c) end of construction, sound rockfill on the inner
shell;
d) ar-cer filling of the reservoir, sound rockfill on
the :'nner shell.
4.2. DAM WITH A SLOPING CORE

Fig. 40 shows the cross section of the Belimo II dam which has
a sloping core. Materials are the same as those used in Belimo
I dam.

! 2000
-~-

o 50 m
I I

CD - CLAY 0- SCHIST AND GREYWAKE {SOUND}

0- SAND AND GRAVEL CD-ROCK {FOUNDATION}

0) - SCHIST AND GREYWAKE {WEATHERED}

Figure 40. Cross section of delimo II Dam.

The vertical and horizontal stresses after filling of the


reservoir (weathered rockfill is in the inner zone of the
shells) are shown in Figs. 41 and 42 respectively. When
comparing Figs. 41 and 36 it can be concluded that, for the
upper zone of the core, the position of the core does not
matter. In the lower zones the stresses (vertical) are higher
on the downstream side for the vertical core and in the
vicinity of the upstream core face the stresses are higher for
the sloping core.
In terms of horizontal stresses the influence of the core
position can be evaluated by an analysis of both Figs. 42 and
39 b. It is apparent that in the upper zone of the core there
is no difference between the horizontal stresses involved but
in the lower part (with the exception of the zone very near
the upstream core face) horizontal stresses are higher in the
vertical core.
It is important to note that at the end of construction,
horizontal stresses in the upstream lower zone of the core are
higher in the sloping core.
The conclusion is that according to the model adopted and
for this type of materials there is no advantage in sloping
424

the core as far as hydraulic fracture is concerned. This is


more or less in accordance with the conclusions presented in
section 3.

1000

1500 2000 2000 2500 2000 1500 1000 500

50 m
1

Figure 41. Vertical stresses in Belimo II Dam after filling of


the reservoir.

1250~ _ _-

250

50m
1===1==1

Figure 42. H~rizontal stresses in Belimo II Dam after filling


of the reservoir.

4.3. INFLUENCE OF THE DAM HEIGHT

All the studies described above were carried out on dams 120
m high. For an ideal frictional material these results could
be used for dams of different height because the distribution
of stresses and strains would be the same only on a different
scale. Concerning stress levels (a measure of the degree of
mobilization of the shear strength, i.e., the ratio between
the installed and the maximum available shear strengths) the
results would be equal independently of the dam height.
Since real materials are very different from these ideal
soils, dam height has an important role in the stress-strain
425

dam behaviour. As an example, Fig. 43 shows the stress levels


at the end of construction for dam heights of 30, 60 and
120 m. As can be observed, there is a significant difference
in stress level values in the three cases. Only the dam with
120 m height shows limited failure zones at the bottvID of the
filters. For the height of 30 m the degree of mobilization of
the shear strength is very low.

0) H .30m

0.2

15m
i==t==:=il

b) H 60 m

0.4

0.4

0.6 0.8 0.95

o 60 m
~

Figure 43. stress levels 6.t the and of construction. Dam


height equai to:
a) 30 m
b) 50 m
c) 120 m
426

Fig. 44 shows the stress levels in the dam with 120 m height
after impounding. When compared with Fig. 4 3c, shows the
interesting change of stress levels in the dam body mainly in
the upstream part where collapse played certainly an important
role.
These results emphasise the difference in the behaviour of
dams of different heights. This analysis deserves more
attention because the occurrence of failure zones by no means
implies less safety regarding ultimate limit states such as
failure due to piping in the core for instance. And it is well
known in the profession that small dams have specific problems
and specific solutions for dealing with such problems.

0.8 0.95

50 m
i=====tl

Figure 44. stress levels after filling of the reservoir for


the dam with 120 m height.

4.4. DEFORMATIONS

As was pointed out in 3.3.4 deformations may be of no use or


even misleading when we try to obtain information about stress
transfer. Nevertheless they are fundamental where analysing
dam serviceability.
Fig. 45 shows the displacements at the end of construction
of Belimo I Dam and Fig. 46 the deformations after the reser-
voir filling. Nearly all the results concerning the upstream
shell are in accordance with experience. Probably in the
downstream shell settlements might be higher. As a matter of
fact, collapse was only simulated (using parameters from the
Beliche dam) in the upstream shell and this is the reason why
larger settlements occurred in that zone owing to submersion.
In the case of Beliche dam and as was already pointed out in
4.1, the action of rain on the downstream shell also origi-
nates an important collapse and consequently large settlements
were observed.
One important type of deformations are the settlements due
427

to creep but forecasts in this area are a difficult matter to


tackle. Empirical relationships such as those of Rodriguez-
-Miranda (1986) can help in fixing the dam camber. It is
assumed that vertical creep deformations are related with the
vertical deformability of the compacted rockfill and the
relationship is based on observed behaviour of fifteen
rockfill embankments (Fig. 47).

o 50 m
I I

Figure 45 _ Displacements on Selimo I Dam a"t the end of


construction,

SCm
t:==t==1 - .1.00 m

Figure 46. Displacements on Belimo I Dam after filling of the


reservoir.

It must be accepted that structural behaviour may in certain


circumstances be dependent on creep, mainly when serviceabili-
ty limit states are being considered. Nevertheless, more
research is needed to attain reliable constitutive equations
that allow mathematical modelling of the phenomenon.
42R

1.8

Q2

10 14

Mv x 101. IkNfm2)

Figure 47. Verticai s~rai~ due to creep after 10 years (6'0)


versus vertical defo:cmability of compacted rockfill (M,,)
(Rodriguez-Miranda, 1986).

4.5. FINAL REMARKS

The procedures just described as an example are not intended


to be used only for the purpose of quantification. The
objective is to use this type of parametric study for a very
instructive comparison between various possible solutions in
such a way that trends recommended by practice may be con-
firmed or adjusted.

5. Safety Evaluation and the Limit States Concept

Ensuring the safety of dams is the main goal of everyone who


designs, builds or operates this type of structure. It is thus
a matter that deserves the best efforts within the profession
and it is evident today that more and more attention is
devoted to safety (risk assessment, safety quantification,
safety and monitoring, standardization and codification,
public opinion, etc).
For common structures the present trend concerning safety is
that deterministic methods for its evaluation are giving way
to semi-probabilistic methods. Nevertheless, since dams are
unique, complex and highly sophisticated structures, important
difficulties may arise with application of the methods
advocated for simple structures.
429

What follows are some brief considerations about safety


quantification of earth-rockfill dams.

5.1. METHODS FOR EVALUATING DAM SAFETY

It is nowadays an unquestionable fact that the classic


stability calculations (limit equilibrium methods) cannot, by
means of the numerical value of a safety factor, quantify the
safety of an embankment dam. Moreover, as regards ensuring the
serviceability of the dam during any significant stage of its
operational life (first filling, operation, rapid drawdown),
such calculations have an extremely limited use.
This does not mean that those methods do not constitute one
of the greatest achievements of soil mechanics. In fact, they
play an important part in the design of embankment dams. They
may, however, be misleading in evaluation of the safety of the
dam if it is considered that the safety factor is in itself a
means of quantifying the overall safety of the dam.
The effort to get round these inconvenient aspects may take
two main courses.
The first consists of considering that equilibrium methods
are useful, but that the numerical values of the safety factor
must have little or no weight in assessing the safety of the
dam as regards a catastrophic failure. Those calculations must
not, therefore, be taken as sUbstitutes for field studies that
are difficult and costly and for the whole investigation
required in order to establish the real character of the
structure concerned. It is also necessary to have qualitative
jUdgements based on experience and detailed studies.
According to Peck (1988), this attitude allows the formation
of a well founded decision for obtaining a safe structure
(safety that is not quantified, but certainly a greater degree
of safety than present knowledge provides). It is, however,
necessary to carry out laborious field investigations, obtain
records with the description of constructional aspects, and,
having access to them, to interpret them, inspect all visible
aspects, locate and interpret all maintenance records and, in
some cases, to install monitoring equipment capable not only
of describing present behaviour but also of forecasting future
behaviour.
It is thus a casuistic procedure consequently based on the
idea that each dam is a case by itself, and must be treated as
such.
The second procedure consists of applying to the ge-
otechnical structures the formulations on safety that for some
decades have been established for structures, particularly
concrete and steel structures, i.e. limit states design.

5.2. BRIEF NOTES ABOUT LIMIT STATES DESIGN

According to Ovesen (1989) it was Brinch Hansen who for the


430

first time used the term "limit design" in a geotechnical


context. The procedure was characterized by the necessity to
carry out in principle, two separate analyses when designing
a structure: one for determining safety against failure and
another for determining the deformations under actual working
conditions. Hansen (1956) linked the limit state design
concept closely to the concept of partial safety factors.
Whenever a structure fails to satisfy one of its performance
criteria it is said to have reached a limit state. The design
consists in considering separately all the possible limit
states (ultimate and serviceability) and ensuring that the
occurrence of each of the limit states is either eliminated or
sufficiently improbable.
The design criterion when analysing an ultimate limit state
(ULS) is simply to design for equilibrium in the design limit
state of failure. Each state implies the consideration of a
failure mechanism. The criterion must be expressed by

where Sd is the design load effect (calculated by using


adequate combinations of design actions) and Rd is the design
resistance effect (calculated on the basis of the design
material parameters). Design actions and design material
parameters are obtained from their characteristic values
through the use of partial safety factors. It is well known
how extremely difficult it is to obtain characteristic values
for geotechnical materials, and the task of quantifying the
partial factors to be used is just now being initiated.
Analysis of serviceability limit states (SLS) implies
different procedures. Fig. 48 highlights the difference
between ULS and SLS. For preventing ULS, a safety margin
(expressed by one or more dimensionless numbers) is intro-
duced. But SLS prevention is attained by introducing a
constraint on the displacement of part of the structure.
The relevant SLS include deformations that may cause loss of
serviceability of the structure, i.e., those which affect its
efficient use or its appearance.
Geotechnical structures pose some intrinsic problems for
utilization of the limit states design concept, but for simple
problems concerning spread foundations, piles, earth-retaining
structures, etc, procedures may even be codified. Neverthe-
less, dams are very complex structures and, in our opinion,
much more research is needed before this type of design can be
adopted as a routine procedure.
The sections that follow will deal with the advantages and
difficulties with the use of the limit states concept in the
earth-rockfill dams.
431

LOA 0

Safety
margin

~serViceability limit state

Allowable deformation

DEFORMATION

Figure 48. VI timate limit s-cates and serviceability limit


states (Ovesen, 1989).

5.3. OVERALL SAFETY FACTOR AND PROBABILITY OF FAILURE

As mentioned above, earth-rockfill dams (as well as embankment


dams in general) may attain limit states that are different
from those resulting from mechanical instability (i.e., those
which it is possible to "calculate"). Nevertheless, it is
essential to carry out mechanical analysis with mathematical
models that can adequately represent the equilibrium condi-
tions of the forces considered.
Embankment dam problems are characterized by the great
uncertainty and scatter in space and time of the properties of
the materials, complex mathematical models, great number of
parameters, etc, but the traditional design procedure is
deterministic. That uncertainty is then taken into account
through a safety factor and its numerical value is obtained
empirically (statistical analysis about the behaviour of dams
already built) but it is evident that this value is only valid
for a certain design methodology (method of calculation,
evaluation of material parameters, etc).
In Fig. 49 it is apparent how the probability of failure may
vary with the coefficient of variation and number of tests for
a given safety factor (F = 1,5).
The great potential of the concept of probability of failure
versus the concept of safety factor (deterministic) is
nevertheless affected by many difficulties such as knowledge
of the acceptable probabilities of failure, numerical evalua-
tion of the uncertainties, role of the probability of extreme
values, etc.
It can be concluded that neither the overall safety factor
nor the probability of failure are without problems when
applied to the stability analysis of embankment dams, the
432

former because it has not the capacity to represent uncer-


tainties in the data, the latter because it is characterized
by a high degree of speculation and is consequently of small
practical interest to the engineer. The best solution is the
parametric study of the limit equilibrium, the great advantage
of this analysis being the adoption of partial safety factors
for the parameters that are really important in the equilibri-
um study, also taking into account their degree of uncertain-
ty.

'">- Safety factor =1.5


~ 20
>- V =0.2
"-
o
a:
w
CD
:>:
::>
z 10

O+-~--~----~I--~--~--~I--~----
10-2 10-4 10-6

PROBABILITY OF FAILURE

a)

'>
z
factor = 1.5
'" Safoty
03

~ 10 tests
ex
~
0'
u.
0

iw
01
u
u.
u.
w
0
u 0
10~2 1cr 6
PRO BABILITY OF FAI LURE

b)

Figure 49. Failure probability for a given safety factor


versus a) the number of -tests; b) the coefficient of variation
(V) (Londe, 1989).

5.4. PARTIAL SAFETY FACTORS AND LIMIT STATES

When designing the dam according to the limit states concept


433

it is necessary to consider:

a) the appropriate limit states (see 5.2);


b) the design situations, i.e. those sets of physical condi-
tions for which it shall be demonstrated that limit states
do not occur;
c) actions and appropriate combinations of them;
d) properties of the materials;
e) partial safety factors affecting actions, material proper-
ties and geometrical data.

The enumeration of all the limit states (ULS and SLS) is a


kind of check-list of all the possible mechanisms of failure
and their characteristics as well as the deformations of the
structure that imply its loss of serviceability. The partial
safety factors will have fixed values (or range of values)
independently of the considered design situations (this is
different from the deterministic procedure of using safety
factor which varies with design situations).
The probability of failure will in principle be independent
of the partial safety factors adopted.
Federico and Musso (1989) pointed out that the definition of
limit states holds for a part or for the whole geotechnical
structure but it does not explicitly take into consideration
the contact (interfaces) between different material so
characteristic and important in structures as earth-rockfill
dams. As a matter of fact there are many phenomena originated
in these zones that cannot be immediately identified as limit
states (for instance, superficial erosion of the downstream
shell material, suffusion in the critical filter, preferential
water paths, damage on the upstream facing, etc) but they may
evolve for a limit state. These occurrences are named critical
events (CE) and due to the complex connections of phenomena
with origin in the interfaces, the consideration of a chain of
critical events, by means of which it is possible to single
out a possible path towards the ULS of the dam, is advocated.
The following chain of critical events are presented as an
example: migration of the core fine particles (CE), internal
collapse of the material (CE), high settlements of the crest
(CE), dam overttoping (SLS ), high pore pressures in the
downstream shell (CE), sliding failure of the downstream dam
shell (ULS). As can be seen, a SLS can be one of the elements
of the chain.
ULS would thus represent the final step of chains of CE so
networks of critical paths representing the scenery of CE that
are to be taken into account in the limit states design must
be envisaged (Bolton, 1989). Dam safety would be assured if
the initial CE of all possible networks are under control.
Nevertheless if initial CE escape full control, or if the
control implies a too high cost, two different strategies are
still possible.
The first one consists in identifying inside the CE chain,
434

at least one CE (CE*) that can be surely controled. The chain


of the evolution of undiserable phenomena towards ULS is
interrupted, but the consequences of the CE preceding CE* must
be accepted and kept under control.
The second strategy is a radical modification of the design
and is the only one valid if each CE of the chain is not
surely controllable. The introduction of new "lines of
defence" advocated by Casagrande (1961) is an exemple of this
procedure and, if we are dealing with interfaces, the pratical
consequence of its application is usually the insersion of a
"transition zone" where a more controllable material can be
used. As a consequence a progressively more complex (but also
more safe) dam structure is obtained.
In Fig. 50 it can be seen that by increasing the number of
interfaces (N) the cost of the dam (C) generally increases as
well as the reliability (R) (the new structure is able to
control a greater number of CE). But it is important to note
the now higher number of limit states to be checked and, above
all, the higher complexity of the geotechnical system dam-
-reservoir-foundation. As a result of this complexity, the
probability of occurrence of SLS (P SLS ) grows and the proba-
bility that the system evolves towards ULS (PULS ) becames
smaller (Fig. 50).

,--------- - ----------,
R

---'
C

Figure 50. Dam safety conditions expressed through the


probability of the ultimate limit states (PULS ) the probabili-
I

ty of serviceability limit states (P SLS ) and reliability (R) -


- and dam cost (C), versus the number of contacts among zones
of different characteristics (Federico and MUSSO, 1989).

The approach proposed by Federico and Musso (op. cit.), on


which the role of interfaces is emphasized, appears thus as a
very promising contribution towards the application of the
limit states on the safety evaluation of earth-rockfill dams.
The limit states method thus emerges as a kind of procedure
that allows a more rational approach to safety analysis.
Nevertheless, it will be shown below that its application to
435

structures of very complex behaviour like earth-rockfill dams,


may raise problems that are likely to be, at least by the
moment, an obstacle to the generalized and practical use of
the method.
5.5. USE OF PARTIAL SAFETY FACTORS IN SAFETY ANALYSIS OF AN
EARTH-ROCKFILL DAM
The problem of the use of partial safety factors for limit
equilibrium analysis has recently been tackled by many authors
(for instance Londe, 1979, 1989; Meyerhof, 1982, 1984,1985;
Nguyen, 1985; Baikie, 1984, 1987; De Beer and Van den Broeck,
1989) .
In all cases the cohesion (c) and the tangent of the angle
of internal friction (tg ) are the material properties that
must be factored by the partial factors (Y M). It is easy to
show that there is an infinite number of combinations of pairs
of values of the 1 M, which, when applied to c and tg ,
correspond to a given value of the overall safety factor (F).
Practical use of the partial safety factors implies a calibra-
tion of the new method against the traditional one in such a
way that the actual and accepted safety level of the structure
is maintained.
As has been previously stated, however, the installation of
a limit equilibrium condition is only one of the many ULS that
may occur in the dam (the particular structure being ana-
lyzed). As an exercise, the safety of an earth-rockfill dam
will then be analyzed for another ULS, e.g., piping in the
core due to hydraulic fracture.
Many factors are determinant for the occurrence of hydraulic
fracture but for the sake of simplicity, only the minimum
principal total stress as one of this main factors, will be
considered. It is also assumed that lower 03 stresses
correspond to less safe situations.
The earth-rockfill dam to be analysed has the geometry of
Belimo I dam (see Fig. 34) but filters are excluded and only
one material was used in the shells. Two models will be
used:linear elastic and non-linear elastic.

5.5.1. Linear Elastic Model. The dam has only two zones
(shells and core) and the elastic constants used will be the
Young modulus (E) and the Poisson coefficient (V). The values
adopted for E and V are:

E, (core) = 180 kN/m2; V, (core) 0.40


E2 (shell) = 340 kN/m2; V2 (shell) 0.20
Fig. 51a shows minimum principal stresses (03) values in the
core at the end of construction (03 values are the data needed
to analyze the conditions favorable to the occurrence of
hydraulic fracture).
436

Now, partial safety factors (Y M ) will be applied to one of


the mechanical characteristics of the materials, for instance
the deformability (E); and as deformability contrast between
core and shells is, in principle, in favour of the hydraulic
fracture occurrence, E, (core deformability) will be divided
by YM and E2 (shell deformability) multiplied by the same
factor.
Figs. 51b and 51c show 03 values in the core at the end of
construction, when YM equal to 1.2 and 1.4 are used, respect-
ively. It can be observed that the most important consequences
are more or less at the core's mid height. But if comparison
is made between 03 at different heights in the center of the
core (see Fig. 52) it can be concluded that 03 values are
higher for YM=I.4 (results for YM=I.2 are quite similar). An
opposite situation is attained if the Poisson ratio (V,) is
divided by y M=I.4(see also Fig. 5Id).

I-----------.l 600 1-------.. 6 00

800 800

a) b)

1-----..\ 6 00

800 1---_-\ 600

c) d)

50 m
t:::===='
Figure 51. Linear elastic materials. Minimum principal
stresses ( 3 ), at the end of construction
a) "1M 1.0 b) y,~ 1.2 (E, and E 2 )

d) Y~I
437

It can be concluded that it is difficult to prescribe an


adequate general use of the partial safety factors for the
mechanical characteristics of the materials.

1/
[ I
Y", --
ELASTIC EL NONLINEAR
t!
I
l
E V K Kur

~=+~
1 1

~- 1,4
1
1
1.4 - - ./
~~~~ I /
-
- - 1 I 1
I -
I - 1.4 1 1,4 ,/'
./ /
// .'
/> /

......
,.'
~.=-
.,'
~
.
....... ;b--
~/ ..
....,...,..-- --
y // //
/' . /

.... ~. ~ -+

v y
~~
50 100
Him)

Figure 52. Minimum principal stresses ( 3 ), along AA, at the


end of construction for different models and partial safety
factors ('V,;).

5.5.2. Non-linear elastic model. A similar exercise was made


assuming materials with non-linear elastic behaviour. Values
of the parameters for the materials of the dit'ferent zones
are those presented in Table 4.1. In the present case the
shells are formed by only one zone of sound schist and
greywacke but filters were considered.
Fig. 53a presents 03 values in the core. Use of partial
safety factors was then adopted in one of the most influent
hyperbolic parameters: K (and obviously Kur ).
The values of 03 in the core for 1M equal to 1.2 and 1.4 are
presented in Fig 53b and 53c, respectively.
Thus in the core

K =
43S

and in the shell

For the filter material the YM value adopted was equal


to 1.
As for linear elastic model is at the mid-height of the core
that some modifications occurred and the 03 values at the
center of that zone, for different heights, were also
included in Fig. 52 (E and V values of the linear model are
equal to Eo and Vo values of the hyperbolic model). The use of
the non-linear model resulted in higher 03 when comparision is
made, at the same level, with those one calculated using the
elastic model referred in 5.5.1. But the adoption of the
partial safety factors also resulted in the growing of 03 as
in the elastic model, i. e., a safer situation is also at-
tained.

600 600
800 800 800
1000 1000 1000

a) b) c)

0 50 m
I I

Figure 53. Non-linear elastic materials. Ivrinimum principal


stresses (03) on the core for

a) b) l' ~1 1.2 c) 1.4

Trying to go a little further in this analysis distribution


of stress levels in the core and filters for the three
situations were calculated and the obtained results are
graphically presented in Fig. 54. It can be observed that the
mobilized shear stresses in the core are higher for YM equal
to 1 which explains the higher values of 03 in the core, for
a certain level, when comparing with the analysis where YM of
1.2 and 1.4 were employed. The use of YM for K values of the
shell makes them stiffer and as a consequence the mobilized
439

shear strengths in the filters becomes higher, as can be


seen in Figs. 54b and 54c.

0.8 08

0) b) c)

o 50m
L=====t

Figure 54. Non-linear elastic materials. stress levels for

a) Yr.\ = 1.0 b) y~\ = 1.2 c)

5.6. FINAL REMARKS

The adoption of partial safety factors in safety analysis of


earth-rockfill dams regarding ULS that can only be attained
when limit equilibrium situations are installed is no matter
of great concern and their use must be advocated.
But when the ULS (or SLS) depend mainly on materials proper-
ties other than the shear strength, use of partial safety
factors is not so straightforward. On the contrary it needs
much more research before this procedure can be used as a
routine. The analyses presented in 5.5 were intended to be
helpful for showing the nature and the difficulty of the
subject.

6. Conclusions

Earth-rockfill dams are safe, economical and versatile


structures being also extremely competitive when compared with
other types of dams.
Recent and important theoretical and practical progresses on
areas tightly linked with these dams resulted in significative
changes in their conception.
Mathematical modelling and the use of adequate constitutive
laws, namely regarding rockfills, is definitely an essential
tool to attain a good and safe design of an earth-rockfill
440

dam. Limit equilibrium analyses don't help very much on this


task.
It is desirable to try to use the limit states concept on
the safety analysis of earth-rockfill dams. Nevertheless much
more research is needed in this area before practical results
can be obtained.

Acknowledgements

The author would like to thank to J. M. Mateus da Silva,


Trainee Research Officer in LNEC, for his valuable coopera-
tion.
441

P..EFERENCES

Adikari, G.S.N. and A.K. Parkin (1982) - "Deformation behav-


iour of Ta1bingo dam". International Journal for Numeri-
cal and Analytical Methods in Geomechanics, vol.6, pp.
353-382.
Arnevik, A., B. Kjaernsli and S. Walbo (1988) - "The Storvatn
dam. A rockfill dam with a central core of asphaltic con-
crete". Norwegian Geotechnical Institute, Oslo, Pub.
N173, pp. 1-8.

Baikie, L.D. (1985). Discussions. Canadian Geotechnical


Journal, 22, pp. 143.

Baikie, L.D. (1987) - "Charts for design and evaluation of


simple earth slopes using total and partial factors of
safety: a review of several available methods". Canadian
Geotechnical Journal, 24, pp. 216-231.

Biarez, J., J.L. Bordes and P. Londe (1970) - "La fissuration


des noyeaux des barrages en enrochements ou en alluvion".
lOth International Congress on Large Dams, Montreal,
vol. 1, pp. 27.

Bolton, M.D. (1989) - "The development of codes of practice


for design". 12th International Conference on Soil Mechan-
ics and Foundation Engineering, Rio de Janeiro, vol. 3,
pp. 2073- -2076.

Campos Pina, J .M. and M.A. Guzman Martinez I (1985) - "EI


Caracol". Behaviour of dams built in Mexico (1974-1984)
Comission Federal de Electricidade, pp. 7.1-7.49.

Casagrande, A. (1950) - "Notes on design of earth dams". Jour-


nal of Boston Society civil Engineers, vol.37, October.

Casagrande, A. (1961) - "Control of seepage through founda-


tions and abutment of dams" Geotechnique, vol. 11, n Q 3,
pp . 161-181.
442

Casagrande, A. (1965) - "Hohe Staudamme (High Dams) ". Mittei-


lungen des Institutes fur Grundbau und Bodenmechanik,
Technische Hochschule Wien, referred by Sherard, J.L., in
II Embankment Dam Engineering", ed. R.C. Hirschfeld and
S.J. Poulos, John Wiley & Sons, New York 1973, pp. 302.

Charles, J.A. (1973) "Correlation between laboratory


behaviour of rockfill and field performance with particu-
lar reference to Scammonden dam". Ph.D. Thesis, Universi-
ty of London.

Clements, R.P. (1984) "Post-construction deformation of


rockfill dams II . Journal of Geotechnical Engineering,
ASCE, vol. 110, N 0 7, July pp. 821-839.

Cooke, J.B. (1984) - "Progress in rockfill dams". Journal of


Geotechnical Engineering, ASCE, vol. 110, No. 10, Octo-
ber, pp. 1383-1414.

Dascal, o. (1987) - "Postconstruction deformation of rockfill


dams". Journal of Geotechnical Engineering, ASCE, vol.
113, NQ1, January, pp. 46-59.

De Beer, E. and M. Van den Broeck (1989) - "Considerations


concerning the safety factors introduced in the stability
calculations of slopes". De Mello Volume, Ed. Edgard
Blucher Ltd, Sao Paulo, pp. 77-88.

Di Biagio, E., F. Myrvoll, T. Valstad and H. Hansteen (1982) -


-IIField instrumentation, observations and performance
evaluations for the Svartevann dam". Norwegian
Geotechnical Institute, Pub. N 142, Oslo, pp. 1-14.

Federico, F. and A. Musso (1989) - "Progetto allo stato limite


di contatti e transizioni nelle dighe di terra II III
Convegno Geosintetici per Ie Construzioni di Terra. Norme
ed Applicazioni. Bologna.

Guillen Bravo, G., S. Uriel Romero and J.R. Perez Rodriguez


(1988) - "Anti-cracking measures in the Canales Dam". Our
work on dam construction. 16th ICOLD - S.Francisco.

Gonzalez Valencia, F. and L. Aguirre Soria (1985) -"Chic-


oasen". Behaviour of dams built in Mexico (1974-1984)
Comission Federal de Electricidade, pp. 6.1-6.34.

Gonzalez Valencia, F. and Gusman Martinez M.A. (1985) - "El


Infiernillo". Behaviour of dams built in Mexico (1974-
1984), Comision Federal de Electricidade, vol. II, pp.
3.1- 3.14.
443

Gonzilez Valencia, F. and F. Saldafia G6mez (1985) - "Belisario


Dominguez (La Angostura)" Behaviour of dams built in
Mexico (1974-1984), Comision Federal de Electricidade,
vol. II, pp. 5.1-5.24.
Hansen, B. (1956) - "Earth pressure calculation". The Danish
Technical Press. The Institution of Danish civil En-
gineers, Copenhagen.
Justo Alpafies, J.L., P. Cafiete Cruz and J. del Campo (1985) -
"EI empleo de rocas de baja resistencia en los espaldones
de presas de materiales sueltos". Revista de Obras Pu-
blicas, Mayo-Junio, pp. 463-471.
Kenney, J.C. and D. Lau (1985) "Internal stability of
granular filters". Canadian Geotechnical Journal, 22, pp.
215-225.
Kjaernsli, B., G. Kvale, J. Lunde and J. Baade-Mathiesen
(1982) - "Design, construction, control and performance
of the Svartevann earth-rockfill dam". Norwegian
Geotechnical Institute, Pub. N 142, Oslo, pp. 1-7.
Knight, D.J., D.J. Naylor and P.D. Davis (1985) - "Stress-
strain behaviour of the Monasavu soft core rockfill dam:
prediction, performance and analysis". 15th Congress on
Large Dams, vol. Q.56, R.68, Lausanne, pp. 1299-1326.
Lafleur, J., J. Mlynarek and A.L. Rollin (1989) - "Filtration
of broadly graded cohesionless soils". Journal of
Geotechnical Engineering, ASCE, vol. 115, N 12,
December, pp. 1747-1768.
Lefebvre, G., J.M. Duncan and E.L. Wilson (1973) "Three
dimensional finite analysis of dams". Journal of Soil
Mechanics and Foundations Division, ASCE, vol. 99 N
SM7, pp. 495-507.

Leonards, G.A. and J. Narain (1963) - "Flexibility of clay and


cracking of earth dams". Journal of Soil Mechanics and
Foundations Division. ASCE, vol. 89 No SM2, pp. 47-98.
Leps, T .M. (1988) - "Rockfill dam performance and remedial
measures" in Advanced Dam Engineering for Design
construction and Rehabilitation, edit. by R.B. Jansen,
Van Nostrand Reinhold, New York.
444

Lloyd, H.E., O.L. Moore and W.F. Getts (1958) - "Rockfill


dams: Cherry Valley central core dam". Journal of the
Power Division, ASCE, vol. 84 No. P04, August 1958,
Part I, pp. 1733.1 - 1733.24.

LNEC (1986a) - "Barragem de Beliche. Observac;:ao da barragem


durante a fase de construc;:ao". Relat6rio 161/86-NF,
Lisboa, Junho.

LNEC (1986b) - "Barragem de Beliche. Colaborac;:ao no controle


de construc;:ao dos aterros". Relat6rio 161/86-NFi Lisboa,
Setembro.

LNEC (1989) - "Barragem de Beliche. Observac;:ao durante a fase


de primeiro enchimento". Relat6rio 6/90-NF, Lisboa, De-
zembro.

Londe, P. (1979) - "R~flexions sur la s~curit~ des barrages".


Annalles des Ponts et Chauss~es, 4 e trim. pp. 37-42.

Londe, P. (1990) - "La s~curit~ des barrages". Revue Franc;:aise


de G~otechnique, NQ 51, pp. 41-49.

Maranha das Neves, E. (1989) - "Analysis of crack erosion in


dam cores. The crack erosion test". De Mello Volume, Ed.
Edgard Blucher Ltda, Sao Paulo, pp. 284-298.

Maranha das Neves, E. and A. Veiga Pinto (1988) - "Modelling


collapse in rockfill dams". Computers and Geotechnics,
vol. 6, N Q 2, pp. 131-153.

Marsal, R.J. and L. Ramirez de Arellano (1967) - "Performance


of El Infiernillo dam". Journal of the Soil Mechanics and
Foundations Division, ASCE, vol. 93, No SM4, July, pp.
265--298.

Mateus da Silva, J.M. (1990) - "Modelac;:ao de descontinuidades


em geotecnia". MSc. Thesis, Universidade Nova de Lisboa.

McCleskey, J. (1988) - Discussion on "Postconstruction de-


formation of rockfill dams". Journal of Geotechnical
Engineering, ASCE, vol. 114, N Q12, Dec., pp. 1453-1456.

McConnell, A.D., J.J. Par~, N.S. Verma and D.A.B. Rattue


(1982) - "Materials and construction methods for the dam
and dyke embankments of the LG-4 project". 14th Congress
on Large Dams, Q.55, R8, vol. IV, Rio de Janeiro, pp.
123-144.
445

Mello, V.F.B. de (1977) - "Reflections on design decisions of


pratical significance to embankment dams". Geotechnique,
N3, September, pp. 281-355.

Mello, V.F.B. de (1983) - "Design trends on large rockfill


dams and purposeful monitoring needs". International
symposium on Field Measurements in Geomechanics,
Zurich, September, pp. 805-826.

Mello, V.F.B. de (1984) - "Behaviour of two big rockfill dams


and design aims". First International Conference on Case
Histories in Geotechnical Engineering, vol. II st. Louis,
pp. 973-947.

Meyerhof, G.G. (1982) - "Limit states design in geotechnical


engineering". structural Safety, 1, pp. 67-71.

Meyerhof, G.G (1984) - "Safety factors and limit states ana-


lysis in geotechnical engineering". Canadian Geotechnical
Journal, 21, pp. 1-7.

Meyerhof, G.G. (1985) - Discussions. Canadian Geotechnical


Journal, 22, pp. 145.

Mikuni, E. (1980a) - "Rockfill dams in Japan". Geotechnical


Engineering, vol.11, pp. 93-133.

Mikuni, E (1980b) - "Dam engineering activities in Japan"


Symposium on Problems and Practice of Dam Engineering,
Ed. by A.S. Balasubramanian, Yudhbir, A. Tomiolo and J.S.
Younger, Bangkok, Dec. pp. 109-130.

Naylor, D. J. (1990) - "Numerical modelling of load transfer in


embankment dams". University College of Swansea, Final
Report, SERC Grant nO GR/E/83887.

Naylor, D.J., E. Maranha das Neves, D. Mattar Jr. and A. Veiga


Pinto (1986) - "Prediction of construction performance of
Beliche dam". Geotechnique, nO 3, September, pp. 359-386.

Nguyen, V.V. (1985) - Discussions. Canadian Geotechnical Jour-


nal, 22, pp. 144-145.

Nonveiller, E. (1988) Discussion on "Postconstruction


deformation of rockfill dams". Journal of Geotechnical
Engineering, AseE, vol 114, nO 12, Dec., pp. 1456-1459.
446

Ovesen, N.K. (1989) - "General Report on Codes and Standarts"


Proceedings 12th International Conference on Soil Me-
chanics and Foundation Engineering, vol. V, Rio de
Janeiro.

Parra, P.C. (1985) - "Previsao e analise do comportamento


tensao-deforma9ao da barragem de Emborca9ao". XVI Semi-
nario Nacional de Grandes Barragens, Brasil.

Peck, R.B. (1988) - "The place of stability calculations in


evaluating the safety of existing embankment dams". civil
Engineering Practice.

Penman, A.D.M. (1983) "Latest geotechnical developments


relating to embankment dams". Ground Engineering, May,
pp. 1927.

Penman, A. D. M. ( 1985 ) - Correspondence. Water Power & Dam


Construction, March, pp. 40-42.
Penman, A.D.M. (1986) - "On the embankment dam". Geotechnique,
vol.36, nO 3, sept. 86, pp. 303-348.

Penman, A.D.M. and J.A. Charles (1973) - "Effect of the po-


sition of the core on the behaviour of two rockfill
dams". 11th International Congress on Large Dams, Q42,
R18, Madrid, pp. 355-339.

Pope, R.J. (1967) - "Evaluation of Cougar Dam embankment per-


formance". Journal of Soil Mechanics and Foundations Di-
vision, ASCE, vol 93, NOSM 4, July 1967, pp. 231-250.

Rodriguez -Miranda, M. (1986) - "Pedraplenes". Simposio sobre


Terraplenes y outros Rellenos, Madrid.

Sherard, J.L. (1953) - "Influence of soil properties in con-


struction methods and performance of homogeneous earth
dams". Technical Memorandum 645, Bureau of Reclamation,
Denver.
Sherard, J.L. (1973) "Embankment dam cracking" in Em-
bankment-Dam Engineering, ed. R.C. Hirschfeld and S.J.
Poulos, John Wiley & Sons, New York, pp. 271-353.
Sherard, J.L. (1984) - "Trends and debatable aspects in em-
bankment dam engineering". Water Power & Dam
Construction, December, pp. 26-32.
Sherard, J. L. (1985) - Correspondence. Water Power & Dam
Construction, Marc~, pp. 40-42.
447

Sherard, J. L. (1986) - IIHydraulic fracturing in embankment


dams" Journal of Geotechnical Engineering, ASCE, vol. 112 ,
nO 10, October, pp. 905-927.
Sherard, J.L. and L.P. Dunnigan (1985) - IIFilters and leakage
control in embankment dams". symposium on Seepaage and
Leakage from Dams and Impoundements, ASCE, Denver, May,
pp. 1-30.
Sherard, J.L., L.P. Dunnigan and J.R. Talbot (1984 a) - "Basic
properties of sand and gravel filters". Journal of
Geotechnical Engineering, ASCE, vol.110, nO 6, June, pp.
684-699.
Sherard, J.L., L.P. Dunnigan and J.R. Talbot (1984 b) - "Fil-
ters for silts and clays". Journal of
GeotechnicalEngineering, ASCE, vol.110, nO 6, June, pp.
701-718.

Sherard, J.L., R.J. Woodward, S.F. Gizienski and W.A.


Cleverger (1963) "Earth and earth-rock dams". John
Wiley & Sons, Inc, New York.

Vaughan, P.R. e H.F. Soares (1982) - "Design of filters for


clay cores of dams". Journal of Geotechnical Engineering
Division, ASCE, vol.108, No Gt1, January pp. 17-31.

Veiga Pinto, A. (1983) - "Previsao do comportamento estrutural


de barragens de enrocamento". Thesis. Laborat6rio Nacio-
nal de Engenharia Civil, Lisboa.

Verma, N.S., J.J. Pare, B. Boncompain, R. Garneau and D.A.B.


Rattue (1985) - "Behaviour of the LG4 main dam". 11th
International Conference on Soil Mechanics and Foundation
Engineering, vol. 4, pp. 2049-2054.

Wilson, S.D. (1973) "Deformation of earth and rockfill


dams" . In Embankment-Dam Engineering, ed. by R. C.
Hirschfeld and S.J. Poulos, John Wiley & Sons, New York,
pp. 365-417.
CHAPTER 15
DYNAMIC BEHAVIOUR OF ROCKFILL DAM
E. YA N AGfSAWA

1. Introduction

It is usually well recognized that earth and rockfill dams can be constructed on even softer
foundation compared with concrete dams. In the earthquake regions there have been
reponed many examples of disastrous raptures of earth dams on soft foundat ion
conditions. Most failures were caused by slope failure induced by seismic ground motion.
The loss of freeboard due to com paction of dam materials and consolidation of the
foundation were also major causes of the overtopping of dams. The possible ways in
which an earthquake may cause the fai lure of an e anh dam proposed by RB .Seed are as
listed in Table l.

Tab] 1 Po si I aus ffailuT fan a h am 197


1. Disruption of dam by major fault movement in foundat ion
2. Loss of freeboard due to differential tectonic ground movement
3. Slope failure induced by ground motion
4. Loss of freeboard due to slope failure or soil compaction
5. Sliding of dam on weak fo undation materials
6. Piping failure through cracks induced by ground motions
7. Overtopping of dam due to seiches in reservoir
8. Overtopping of dam d ue to slides or rockfalls into reservoir
9. Failure of sE.!I.I ~ or outlet wo~k_~ __ ~~__ ~~__ " _ "..

Tecton ic ground movement and fault movement are of course problems of prim ary
importance for dam e ngineering. Piping failures and o vertopping due to seiches in
reservoir are al so of great importance in evaluating the safety of earth and rockfill dam s.
However in design practices we usually focus our attention on the problems of the stability
of slopes during severe vibrations. In this text we wDuld like to limit our problems to those
which concern the effec ts of vibration on a dam body, and we will not go further into the
problems of seism ic source mechanism and input earthquake waves.
The response of fil l dams to incident seism ic waves will be affected by the shape of
valley, the stiffness of foundation and mechanical characteristics of embankment materials.
Although there are a lot of examples of failure of earlh dams during eanhq uakes,

.,
fo rtunatel y very few cases of seve re damage to roe kfill dams have been

E. M/Jranlw Jas N~u$ (~d.), Ad,'{JllUs ill Ra<:kJiIl Slr"CI"r~$. 449-470.


10 1991 KI" ....~r Academic Puhlishers.
450

reported so far, and therefore it is already well recognized that rockfiII dam is
more resistant to earthquake than earth dam. However there are a lot of arguments
concerning the safety of rockfill dams against earthquakes, since the failure of a high dam
is undoubtedly a catastrophe to downstream inhabitants. In order to improve the
construction technique of earthquake resistant structures and to enhance earthquake
resistant design methods, it is necessary to determine the mechanical properties of fill
materials and investigate the seismic response behavior of rockfill dams.

2. Damage to Fill Dams due to Strong Earthquakes


2.1 DAMAGE TO FILL DAMS BY STRONG SHAKING

In the seismic regions there were a lot of serious cases of failure of soil structures and
natural slopes due to strong shakings. Most soil structures such as levees, reclamation
dikes, and highway embankments often suffered severe damage because of strong
earthquakes. Slides in slopes and cracks in the crest are often found in embankments which
have been built on relatively weak foundations. Concerning to earth and rockfill dams
there have been reported a large number of small earth dams damaged during strong
shakings. During The Oga Earthquake of 1939, more than 50 small irrigation dams were
reported to have been damaged by the earthquake and 12 cases among them were complete
dam failures (Akiba, 1939). Most of the failures occurred a few hours or up to one day
after the earthquake. During The Tokachioki Earthquake of 1968, 29 small earth dams
among 94 dams for irrigation purposes were damaged in Aomori Prefecture and 7 among
them were destroyed by the shaking. the damaged dams are concentrated in a region where
strong rainfall had been observed prior to the earthquake (Shibata et aI, 1971). This rainfall
apparently amplified the damage to soil structures.
There are some examples of earthquake damage to well constructed rolled fill dams in
Japan. Ainono Dam, a uniform type earth dam with a height of 40.8m, was shaken by a
local earthquake with a magnitude of 5.8 in 1973. Longitudinal cracks with a width of 8 to
25cm were found on the crest, the backfill of a stone wall on the crest subsided by 25cm
and a lighting pole was inclined by the subsidence. The Makio Rockfill Dam, a center core
type dam of 104.5m in height and 260.0m in length, suffered damage in the Naganokenn-
Seibu Earthquake of a magnitude 6.8, September 14, 1984. Longitudinal cracks were seen
on the crest up to the depth of 1.5m and also on the down stream slope cracks and a
settlement of 30 to 40cm were seen. Strong motion seismometers were installed in the
dam, but unfortunately due to stoppage of electric power seismic records could not be
obtained.
In Table 2 the maximum acceleration values ever observed at fill dams are listed. The
Morgan Hill Earthquake shook Coyote Dam in 1984. The maximum acceleration recorded
on the crest was 1.29g in transverse direction, 0.9g in longitudinal direction and O.4g in
updown direction. Despite the large acceleration only slight damage was found in the dam
body. According to the investigation made after the earthquake, three major cracks were
found in the upstream slope, along the crest and near the concrete spillway section. The
crest subsided by 6.7cm and was displaced horizontally by 3.7cm in the downstream
direction. These cracks seemed to have been formed by the displacement of the surface
rockfill mass of the slope. After a careful investigation it was concluded that the cracks do
not have any significant effect on the safety of the dam.
During the Mexico Earthquake of 1985, two major rockfill dams were reported to have
been damaged. La Villita Dam, a center core type rockfill dam, with a height of 59.7m and
a length of 420m was shaken so much that it cracked on the crest. The maximum
451

acceleration observed at the crest of the dam was 0.45g in transverse direction. The dam
crest subsided by Ilcm at the middle point of the dam axis and horizontal displacement
was also found to be l5.4cm in the upstream direction. Two major continuous cracks were
seen intermittently on the both sides of the crest. The upsream side crack has a length of
80m and width of 0.5 to Scm, whereas the downstream crack was 130m long and 5 to
l5cm wide. The cracks are located on the extension of the core slope lines. El Infiemillo
Dam is a 148m high rockfill dam, which has steep slopes with a ratio of 1:1.75. There was
found approximately IOem of subsidence in the middle point of the crest and also two lines
of long cracks were seen on both sides of the crest, which may affect the safety of the core
and stability of the dam.

Table 2. Damage to earth and rockfill dams during earthquakes


Dam Type reak Acre I (g) [pIC Dist Earthquake Damage

TarumlZu 036 TPlest ) 40km Off Mlyagl


(,ock fill Cracks
II o 43m 03)U4(tstl 1978

037 TR (racks
Nagar a 035 LU
off~Chlba
HOrl? Disp 2 Orr
Earth 29km
II o 52m 1987 SubSidence 3 Om
033 LJD

Naganurr; a [acth
o I ~ TR Local
lI o Sm Test rill
o 13 LN 20km No Damage
023 UD 1988

Cracks,
[iinfiernilio 034 TR MeXICO
W2~ 150mm
Rock[ III 62kC"
H= 146m (Middle of Slope) 1985
SubSidence 11.3cm

La V 1IIIta Cracks,W5-150mm
Mexico
Rockflll 0,45 TR - Subsidence 20cm,
H=60m 1985
Parapet Wa II

I 29 TR Cracks, W30mm ,L3m


Coyote Morgan Hili
Rockflll 0,90 LN 25km Horlz Disp 74cm
H-3Sm 1964
047 UD SubSidence 67cm

0,45 TR Cracks,
0,34 LN Lorna Prieta Horlz, DISP 7,4cm
Lex Inton Earth 6km
0,20 UD 1969 Subsidence 27cm

17km Morgan Hili


0,62 TR Shallow Cracks
1964
Anderson Earth
043 TR
0,32 LN Loma Prieta Cracks, W4Srrm
28km
0,23 UD 1989 1I0"z.Dlsp 4 Scm

Recent cases of damage to earth and rockfill dams due to an earthquake took place
during the Lorna Prieta Earthquake of 1989. Eight earth dams and a rockfill dam were
reported to have suffered cracks and subsidence in the embankment bodies. These dams
are located within 30km from the epicenter and fairly large accelerations were observed at
the foundation and on the crest of two dams among them. Rockfill dams seem to be
resistant to considerably high acceleration, while earth dams are not always as durable as
rockfill dams.
2.2 SOIL LIQUEFACTION

Soil liquefaction is one of the major items which will cause the serious damage to the
foundation of the dam and the dam body itself. When very loose sandy soil is in a
452

submerged or fully saturated state and strong shaking or a vigorous shock is applied, the
soil will lose shear resistance and the soil mass will behave as a heavy viscous liquid. The
loss of shear resistance is caused by an increase in pore water pressure generated by
repeated loading during an earthquake. The excess pore water pressure reaches to the
overburden pressure when the soil is completely liquefied by the shaking. Liquefaction of
sandy sediments during earthquakes is often recognized by the phenomena of sand blow
or mud blow where small sand volcanoes are observed on the ground surface. Unequal
subsidence and lateral displacement often accompanied by wide cracks or opening in the
ground are seen when the sandy ground is liquefied. Soil structures and slopes subside and
spread laterally by a large amount of displacement which is induced by the increase in pore
water pressure.
The most famous slope failure of a dam due to liquefaction is the case of the Lower San
Fernando Dam during the San Fernando Earthquake of 1976. A major slide occur in the
upstream shell of the dam and only 4 to 5 ft of badly cracked material remained above the
reservoir water level. Eighty thousand people living downstream of the dam were
immediately ordered to evacuate and steps were taken to lower the water level as rapidly as
possible. The factor of safety obtained by the pseudo-static approach was found to be 1.3
using a seismic coefficient of 0.15. A dynamic response analysis was made using a time
history of acceleration which was determined from the seismoscope record recovered from
the slide debris (Seed et aI, 1975). The result of the analysis indicates that a larger part of
the zone of hydraulic fill sand was liquefied during the earthquake and the factor of safety
was reduced to 0.8 a short time after the earthquake. The process of the failure was clearly
described by this marvelous study.

3. Dynamic Properties of Fill Dam Materials


3.1 DYANAMIC DEFORMATION CHARACfERISTICS OF FILL DAM
MATERIALS

It is necessary to evaluate elastic moduli and damping properties of soils and rocks in order
to analyze the earthquake response of earth and rockfill dams. In the field of geotechnical
earthquake engineering it is already well established that the responses are usually
calculated in the case of shear wave incidence. By introducing the complex shear modulus,
material damping can be automatically taken into account when the equation of motion is
solved either analytically or numerically. In this case damping may be frequency dependent
when viscous damping is considered and it may be frequency independent when usual soil
material is considered. Shear moduli and damping factors are the most important
parameters for the evaluation of dynamic responses of soils. Deformation characteristics of
soils show substantial non linearity in the stress strain relationship. Therefore the secant
modulus of the stress strain curve will gradually decrease as strain increases. Energy
dissipation during one cycle of loading will increase as the strain amplitude increases. Then
the damping factor must be also strain dependent. These dynamic properties of soils can be
determined by in situ tests or by laboratory tests. Through in situ measurement of shear
wave velocity by elastic wave velocity logging or field array observation the shear modulus
of soils can be estimated. In laboratory tests, dynamic triaxial tests, dynamic simple shear
test and resonant column test are used to determine the shear modulus of soils.
The resonant column method was first employed by !ida (1938) to study wave
propagation in vertical columns. Since that time many papers have been published
concerning the laboratory measurement of shear wave velocity. Hardin and Black (1968)
453

have indicated the various quantities which exert an influence on the shear modulus of
soils.
G = f(a o', e, H, S, 'to' C, A, f, t, h, T) (1)
where CJo': effective octahedral normal stress, e: void ratio, H: ambient stress history
S : degree of saturation, 'to: octahedral shear stress, C: grain characteristics, A: amplitude
of strain, f: frequency of vibration, t: time, h: soil structure and T: temperature including
freezing.
In the resonant column test a hollow cylindrical specimen was used to make stress
distribution as uniform as possible. It was found that a solid cylindrical specimen can be
also used since differences of both test results are negligible. Strain amplitude which is
available in the resonant column test ranges from ten to -6 to -3. In cases of strong
earthquakes strains in the ground exceeds this range of strain amplitude. For larger strain
amplitude dynamic triaxial tests shall be applied, in which strain amplitude ranging from
ten to -4 to -1 can be easily attained. The simple shear test is sometimes used to determine
the shear modulus of larger grain sizes. Using these equipments the dynamic shear
modulus and the damping ratio can be determined in terms of shear strain amplitude.

Table 3. Shear modulus of soils and rockfills


Clean Sand
(2.17 - e)'
Hardin & Richart G ~ 700 a 0>

(1963) o 1+ e m

(2.17 - e)' 0 >


G 0 =330 a.
I+e
(2.17 - e)'
Tatsuoka et al G 0 =900
1+e
am O.lI
(1977)
(2.17 _ e)'
Kokusho (1980 ) G 0.'
= 840 o.
0 I+e
Clay

G = 330 (2.973 - e)'


Hardin & Black
o I +e
am OJ
(1968)
(2.973 _ e)'
Zen et al (1978 ) G 0.'
=200 - 400 am
0 I+e
(7.32 _ e)' o.
Kokusho et a1 G =9
0 I+e
am
(1982)

Rockfill Materials
,
Baba (1980 ) G ~ 440 (2.97 - e) a ma "
a 1+ e
(2.17 - e)'
Sasaki (1989 ) G ~ 1100 a m D4
o 1+ e
Core Materials
(2.97 _ e)'
Baba et al (1980 ) 395 a."
G
a
~
I+e
a.

It is obvious that the shear modulus decreases and the damping ratio increases as strain
amplitude increases. This fact is well known as the strain dependency of the material
454

constants. Go or the shear modulus at small amplitude can be expressed by a function of


void ratio e and effective confining pressure <Jo.
Go = A f(e)<Jo'm (2)
for soils only if effective mean stress and shear amplitude are defined. In Table 3 the
empirical equations for the estimation of the shear moduls Go of soils and rockfill materials
are listed.

3.2 DYNAMIC STRENGTH OF SOILS

It is said that Collin has first found that shear strength increases when a soil sample is
sheared very rapidly by using double shear apparatus. After some decades A. Casagrande
opened a new horizon in studying the dynamic shear strength of soils. He studied the effect
of loading rate on the shear strength of soils and found that the shear strength of
Cambridge Clay and Manchester Sand exhibit distinct increases when the loading rate is
increased. D.W.Taylor and R.V.Whitman studied simultaneously the effect of the strain
rate. According to the test results of R.V.Whitman shear strength was increased by 1.1 to
1.2 for sand and 1.3 to 2.0 for cohesive soils. B.B.Shimming performed strain controlled
shear tests using a special direct shear apparatus by which a soil specimen can be sheared
in up to 10 to 30 mili seconds. From the test results he concluded that in the case of dry
sand there is no strain effect at all neither in a loose state and a dense state. For cohesive
soils, his experimental results also showed that the strain rate effect depends on normal
stress and the maximum value of strain rate effect will be 2 for clay. These results are listed
in Table 4. It may be concluded from the table that for cohesive soils we can expect strain a
rate effect from 1.2 to 2.0 and for sandy soils 1.0 to 1.2.
The situation is completely different when cyclic loading is applied to saturated soils.
H.B.Seed performed cyclic loading tests on cohesive soils and simulated seismic loading
conditions by adding pulsating load to dead load (Seed et aI, 1966). He drew some lines
of the number of cycles by which the sample fails because of the pulsating load. The
dynamic strength is designated by a ratio of the pulsating load to the static strength, and
therefore the cyclic strength which failed in one cycle is larger than one hundred percent
because of the strain rate effect. It might be easily understood that cyclic strength is usually
larger for clayey soils. However sandy soils fails sometimes under smaller cyclic load than
static. This phenomenon is recognized as liquefaction. The liquefaction strength of sandy
soils can be obtained by cyclic simple shear tests or cyclic triaxial tests. If the stress ratio
between dynamic shear stress and normal stress or the ratio of deviatoric stress and
confining stress are taken as the ordinate and the number of cycles which is needed to
liquefy are taken as the abscissa in a logarithmic scale then the liquefaction strength curve
can be drawn. Cyclic stress, strain and pore pressure are measured during the test to be
able to draw an effective stress path on p-q plane or <J-'t plane. These undrained shear
behaviors must be simulated by constitutive models in numerical analyses.
Liquefaction strength is often referred to as the relative density of sand. For the loose
state of clean sand the linear relation between Dr and 't/<J can be seen. Dense sand has
larger strength compared with loose sand. The pore pressure buildup can be seen also in a
dense sand during small shear strain. However as soon as the shear strain exceeds a certain
value, the soil becomes dilative causing a decrease in the pore water pressure. As a result,
the condition of zero effective stress occurs only momentarily, and the soil retains
considerable shear strength even after the liquefaction. Well compacted rockfill materials
exhibit considerably greater shear resistance than dense sands under cyclic loading
conditions (Matsumoto et aI, 1987). High permeability of the materials will prevent a full
455

development of pore pressure and reduce the duration of fully liquefied condition during
the period of earthquake shaking (Wong et al, 1975). The possibility of liquefaction of
rockfills is likely to be extremely rare even in highly seismic areas.

Table 4. Strain rate effect of soils


Dynamic T esls for Clay
Tt'st Pevice Cuntlnill'j Time to St ra lil t{dte
YEAR Author
& Cundition Pressure Failure Effect
H~~lTldrk!-'

Triaxial 4 min
1944 Taylor Undrained 1.15
-8 day
Unconfined 0.02 s
A.Casagrande 3-6
1948
& Shannon Undrained
Tr iaxi a 1
kgf Icm 2 -1000 5
1.4-2.6 Prismatic

A.Casagrande 0.01 5
1948 Unconfined -240 s 2.4
& Shannon

A.Casagrande Triaxial 3-6 0.05 s


1948 Shannon kgf/cm2 -500 s 1.5
& Undrained
A.Casagrande 0.01 s
1951 2.0
& Wilson -1 month
Unconfined 2.1, 3.0
1953 Taylor/Whitman 0.005 5
& others
& Triaxial 6.0 1.3-2.0
1954 Undrained kaf/em 2 -300 s
Unconfined
Whitman 0.05 S 1.3-2.0
1957 & Triaxial
Undrained -300 s
Triaxial 0.0025 S
1961 Richardson 4.2 kgf/cm 2 1.6
Undrained -300 s
Unconfined 1.57
Whitman 0.0015 s
1962 & Triaxial 4.2 kgf/cm 2 -1.71
& Richardson Undrained -300 S
0.01 s
1966 Shimming Direct Shear -50 s 1.6-2.0

Dynamic Tests for Sand


Test Device Confining Time to Strain Rate
YEAR Author Remarks
& Condition Pressure Failure Effect
Triaxial 15 s
1947 Taylor Vacuum 1.0
-300 s
A.Casagrande Triaxial 0.3-0.9 0.03 5 1.1
1948
& Shannon Vacuum kgf/cm -2100 s
Seed Triaxial 0.02 s 1.15
1954 & Lundgen Vacuum 600-900 s -1.20
1953 Taylor, Whitman Triaxial 0.3-1.0
Vacuum 0.005-3005 1.1
1954 & others & Undrained kaf/cm2 0.2-180 s 2.0
Triaxial Dry Sand
Whitman Vacuum 4.2 kgf/em 0.005 s 1.0
1957 Satul~t;~~
& Undrained -300 s 1.1-2.0
Whitman Triaxial
Vacuum 0.005 s l.0
1960 & Heady & Undrained -300 s 1.4-2.0
Whitman Triaxial 0.025 s Satulated
1962 l.4
& Hear1y Vacuum -5 S looseSand
Triaxial 0.4 0.7 0.013 s Silty
1962 Heady 1.4 1.1-1.2
Vacuum , k~f~cm2 -4 s Sand
0.01 s
1966 Shimming Direct Shear l.0
-30 s

Liquefaction of fine containing sand is now the subject of many current studies which
are done for the purpose of analyzing the stability of hydraulic fill or tailing dams. Recent
studies are directed at the effect of fine content, three dimensional effect and the rotation of
the principal stress axis.
456

4. Response Analysis of Fill Dams


4.1 SHEAR BEAM THEORY

The response of an embankment to earthquake excitation is generally affected strongly by


the soil conditions underneath the dam body. In order to verify the effect of surface layers
on the response characteristics of an embankment, a triangular embankment body on a
horizontal surface layer was considered as shown in Figure 1. Considering two waves in
the layer, vertically ascending and descending, respectively, displacement at any point in
the system can be described as a function of time t and the depth z. If the phase differences
of the waves are taken to be zero at the ground surface (z=H), then the shear wave at the
bottom of the embankment can be written as Uocos(pt). Hence the relative displacement of
the embankment body. ~ is obtained by the solution of forced vibration of a shear beam,
; = L Fj(z)Q j(z) (3)

where (PZJ = J o(z.Z)


Fj(z) = J 0 V l 3 H (4)
2

~i(~)UO
Qj(z) = 21 2 {Djcos(Pt)+ Ejsin(pt)} (5)
D j +Ej
in which Pi=ziV 3/H; Ili=the participation factor of i-th order; Jo = Bessel function of zero
order; zi= the i-th root of Jo; V3 = the shear wave velocity of the embankment material.

Embankment
~-.---- ---I
03 V3 G3

Surface Layer
h
P2 V 2 G2

Figure 1. An embankment on a surface layer

By equating displacement and shear stress at each boundary of the layers, amplitude and
phase difference at any point of the system can be solved in term of sinusoidal input wave
amplitude AI. The ratio of response amplitude Ac at the crest of the embankment to input
wave amplitude Al gives the magnification factor of the system and the ratio can be written
in the form
Ac 2 j(1+ L,CjD/ + (L,CjE/
(6)
AI ~
where B is a function describing the effect of the foundation vibration on the embankment
vibration and
457

(7)
where K J and K2 are parameters expressing the effect of the embankment on the
vibrational characteristics of the ground. In the Eq.(7),
ri=cos(phiNi), si=sin(phiNi), ~=Pi+JVi+/PiVi lli=2/zjJ 1(zj),
Di=1-(p/Pi)2, Ei=TlP/G, Ci=(P/Pi)2/(Di2 +Ei2),
and Kl=~LqEi(p/Pi)Jl(zi) (8)
K2=u2LCi(Di+Ei2)(p/Pi) J1(zi) (9)
in which r and V represent the density and the shear wave velocity, respectively, and
subscripts 1, 2 and 3 designate the materials of the bedrock, the surface layer and the
embankment, respectively. The response of the embankment can be described by a
function of height, rigidity and viscous damping of the embankment, and thickness and
rigidity of the surface layer.
In order to investigate the effect of ground conditions on the vibrational characteristics
of an embankment, the simple case of a two layered foundation was studied. The response
at the crest of the embankment to incident S-waves was calculated for various constants
and dimension of the foundation layer. In this case viscous damping was considered only
in the embankment to be 5%. Figure 2 indicates response curves of the embankment on
layered foundation. The broken line indicates the response curve of the triangular
embankment body, whose amplification ratio was calculated by dividing the response
amplitude at the crest by the input wave amplitude at the bottom of the embankment. The
dotted line in the figure indicates the response curve of the layered foundation without any
20 -- __ Embankment
........... Ground
I-.
0 - - Embankment Ground
+J
U
ru
~

c
0 10
.,-<
+J
ru
u
.,-<
4-<
.,-<
M
Q,

------- ---,..
E
<l;
0~--------~5~---------1~O~--------~15
Frequency (Hz)
Figure 2. Interaction between embankment and foundation

effect of the embankment. It can be seen from these response curves, that the natural
frequency of the embankment body is calculated to be 8.5Hz and that of the layered
foundation is 7.8Hz. The natural frequency of the whole system designated by the solid
line is 4.8Hz as shown in the same figure. It should be noted from these results that the
resonant frequency of a dam will be affected by foundation condition and eigen values can
not obtained from mere superposition of the individual vibration system.
458

20

c
o
~10
u
~

O~------~5------~10~----~15
Frequency (Hz)
Figure 3. Energy dissipation through the foundation

It is well known in the case of layered foundation that the response of the embankment
is apparently decreased by wave dissipation and we can regard the energy dissipation as
radiation damping in addition to the viscous damping of the system. Figure 3 indicates the
response of the embankment in respect to a different impedance ratio ul, It can be seen in
this figure that the response curves can be regarded as if the damping ratio was increased
when the impedance ratio was increased. This is clear evidence of radiation damping which
can be assumed during two dimensional analyses of fill dam section. For dynamic
nonlinear response analyses, radiation damping can be often taken from 5% to 10%
according to the foundation conditions. It is also pointed out that to simulate the observed
behaviors of actual dams during earthquakes a larger value of damping ratio must be
assumed in addition to the material damping.

4.2 RESPONSE ANALYSES OF ROCKFILL DAM

The simplest ways of dynamic response analysis might be linear elastic assumption.
Linear elastic finite element analysis can be made by dividing a region of continuous
materials into numbers of triangular or square elements. By defining displacement vectors
{V} on nodal points, the equatiol). of motion can be written as follows.
[M](V} + [C]{V} + [K]{V} = (Pd (10)
The damping of the system can be evaluated by material damping and radiation damping.
Material damping depends on the soil characteristics, stress state and strain state. Material
damping is usually recognized as being frequency dependent and the empirical relations
obtained from experiments are often used to evaluate the values of the damping ratio.
Damping is usually regarded as viscous damping and so called Rayleigh Damping [C) =
arM] + b[K], for example, can be used for the analysis. In this case damping ratio plays a
significant roll in the evaluation of maximum response acceleration. As we have discussed
in the previous section, damping of the system can be evaluated by material damping and
radiation damping. Radiation damping depends on the foundation conditions. The
magnitude of radiation damping sometimes can be decided to be 5 to 10% by engineer's
judgement.
Equivalent linear analysis is a practical calculating technique in which the non-linearity
of materials can be taken into account. The procedure is primarily based on linear elastic
analysis however the elastic modulus of each element is changed in the next calculation in
proportion to the maximum strain which can be given from the time history of the element
in the previous calculation. Then new stiffness matrix is formulated using the new elastic
modulus and the elastic response will be again calculated. For the evaluation of the elastic
modulus, strain dependency of the shear modulus and the damping ratio is taken into
459

consideration. Repeating the above mentioned procedures, the maximum acceleration of


each nodal point converges to a value if reasonable ranges of errors are allowed. Although
the procedure is rather simple compared with nonlinear analysis, the result can be realistic
only if adequate parameters are given for the analysis. To get a more accurate
approximation it would be necessary to accumulate the case histories of the response of the
embankment dam to compare the calculation results with observed ones.

~ Elm
70,
i
50'
I
3O! s

10i

Figure 4. Plan and maximum cross section of Tarumizu rockfill dam


1~=
__ '~"'t-~:O.262sec 3~16sec
~- 4'~D
' -' ' " ~- -=z~'-- ~
~-I

~41'
2nd~sec
:\', ~4sec
--~ ""~-' .- -~

Figure 5. Vibrational modes and natural frequencies

Response analyses of rockfill dams are often made by two dimensional finite element
models of the maximum cross section of the dam in plane strain condition. This is valid
only for the long dams in the axial direction. However this assumption can not be applied
to the dams which are constructed on a narrow valley because they will show effects of
constraint by the side abutment. Accordingly for the case of The Tarumizu Rockfill Dam, a
three dimensional finite element model was considered. The model consist of 104 nodal
points, 321 tetrahedral elements and 66 frustum elements. Young's modulus of core
material and rockfill was so decided that the natural frequency of the first order coincide
460

with the observed value of 3.8Hz. Namely Young's modulus of the core material was
assumed as 50,OOOtf/m3 and that of rockfill materials as 90,OOOtf/m3. Figure 6 indicates
some vibrational modes obtained by this analysis. The first mode is a horizontal
symmetrical mode in x-direction (transverse direction) with the natural frequency of 3.8Hz.
The second mode is also a horizontal symmetrical one in y-direction (longitudinal
direction). The third is horizontal anti-symmetrical and the fourth is the vertical vibmtion
mode of the natural frequency of 5.1 Hz. The natural frequencies obtained by the
calculation coincide well with the observed values.
In order to simulate the response of the dam, a preliminary investigation was made, in
which the acceleration records observed at the base of the dam were used as the input
waves to the model. In this case the damping factor was assumed to be 0.15 to get a good
approximation of the response. This value of damping is fairly large in comparison with
the estimated value of damping from the amplification curves. By using these material
constants, nonlinear response analysis was made for the input wave, which was observed
in the gallery of the dam during the Miyagikenoki Earthquake of 1978. The maximum
acceleration of the input wave in the transverse direction was 180.6gal and 234.9 in the
longitudinal direction. Figure 6 indicate the calculation results of linear and nonlinear
analysis. the maximum acceleration of the nonlinear response was 354.0gal and this is
smaller than that of the linear analysis by 7.4%. The magnification factor is 1.96 for the
nonlinear analysis.

400 Linear
Nonlinear
r-< 300
'"
lJ'
200
c
o
:;:1 100
...'"
~ O~~~~~~~~~
QJ
lJ
~-100

-200

-300

Figure 6. Linear and nonlinear responses of the dam

In this study, only horizontal movement of the foundation is considered. For three
dimensional analysis, it should be analyzed by three dimensional input waves. The result
will change when the phase difference between left and right abutment is considered. To
simulate the actual behavior of the dam more precise investigation of the seismic response
of the existing dams will be needed.
In the above discussion total stress was considered to describe the stress strain relation of
any soil materials. It is well recognized, however, for the soils submerged by reservoir
water or saturated by seepage flow, pore pressure plays a significant roll in the stress strain
relationship of the material. For example the shear modulus will decrease if pore pressure
is increased by strong shaking. Fortunately we have already enough knowledge to evaluate
the pore pressure generated by dynamic loading, we can easily estimate effective stress in
an element. The way of obtaining the effective stress is related to the theory of plasticity.
Constitutive equations based on plastic potential are often used to obtain plastic shear strain
as well as plastic volumetric strain. Assuming an undrained condition, induced pore
461

pressure can be calculated from the volumetric strain in incremental form, which will take
place when drainage of pore water from the element is allowed.

5. Liquefaction Analysis
5.1 THE CONSTITUTIVE EQUATION FOR UNDRAINED SHEAR BEHAVIOR OF
SANDS

There has been a large number of papers published on the undrained behaviors of sand
under cyclic loading conditions, and great deal of effort has been devoted to constitutive
formulation of the behavior. In these studies constitutive models based on the elasto-plastic
theory are often used to simulate the cyclic undrained behavior of sands. The liquefaction
of sands is a well known behavior resulting from the loss of effective strength of soils due
to generation of pore pressure during cyclic loading. A sudden increase of deformation will
appear when pore pressure reaches confining pressure or overburden pressure. This causes
unequal subsidence, large deformation and loss of bearing capacity, which will result in
damage to structures and foundations. The anisotropic hardening model would be one of
the typical models applicable to evaluate the accumulation of pore pressure under seismic
loading conditioned. Ghaboussi and Momen(l982), Mroz and Norris(1982) and
Poorooshasb and Pietruszczak(l986) would be typical examples of the anisotropic
hardening model. A simple anisotropic hardening model proposed here is based on the two
surface model proposed by Dafalias and Popov (1976). Special attention was paid to
volumetric strain behavior during unloading and reloading.
We define an isotropic hardening surface in stress space;
F(CJij'A,) = 0 (11)
The flow rule for the virgin loading is given by the following equation;
e/= ~(aa:ij)~ and kkP=n(AI)~1 (12)
where eij denotes the strain deviator and 0' denotes the deviatoric part of the enclosed
quantity. neAt) is dilatation factor which defines the relation of the dilatancy rate with
respect to equivalent shear strain rate At. For the isotropic loading surface we use
F = g - R(S)k(A1)p' =0 (3)
where 1'2 denotes the second invariant of the stress deviator; namely 1'2 and Sij ::CJij-
0/3)CJkk<:\j' R(S) denotes the effect of the intermediate principal stress on the shape of
the isotropic loading surface, and p' denotes the effective mean stress. For the function
form of k(At) we employ the following hyperbolic relation:
A
k(~)- (14)
a + bA
1
11.1 -
1
For the cyclic behavior of sands within the isotropic hardening surface, we use the
following loading function
f = j31~ - R(8)k(A )p' = 0
1 (15)

where J* 2=(l/2)(s*ijs*ij) and s*ij=sij-Uij,


462

and Qij denotes the back stress tensor indicating the position of the center of the loading
surface and GoO"s) is the isoropic hardening parameter indicating the degree of isotropic
hardening during the cyclic deformation of sands. The schematic representation of the
loading surface in principal stress space is shown in Figure 7. GoO"s) was assumed to have
a functional form
Go(A s)=b(1- exp(aoAs)) (16)
The following flow rule is assumed to derive the plastic strain deviator;

ei/ = ~ (a~fiJf (17)


where 0' denotes the deviatoric components of the gradient of the loading surface for
cyclic loading with respect to corresponding stress components.

Phase TransformatIon
Bounding

02
ILoading
Events

I Phase Transformation
I
Bounding

"t
I
l

I
Figure 7. Loading surfaces on I
I
I
the deviatoric plane : As ... ~
/ I
I ... '~ ,~1/~' I
It would be necessary to describe the t I" ,,, I I"

Loading Events
difference between the role of the
parameters of At and As. Figure 8 shows Figure 8. Explanation of hardening parameters
the schematic representation of the
parameters At and As under typical loading events. The parameter At is a monotonic
increasing function which operates for virgin loading only. On the other hand As operates
in both virgin loading and cyclic loading within the equivalent plastic modulus surface. As
will be set equal to zero when stress reversal occurs in the stress state beyond the phase
transformation lines, which implies the shrinkage of the loading surface to a point. Thus
the parameter At has the role of memory regarding the previous maximum shear in terms
of the stress ratio, and the parameter A.s plays the role of defining the gradual increase of
shear resistance when sand is subjected to repeated low shear stress and the drastic change
of volumetric deformation behavior during the stress reversal beyond the phase
transformation line.
463

5.2 PORE PRESSURE GENERATED DURING EARTIIQUAKE

After rewriting the constitutive equations into convenient stress and strain quantities for the
triaxial and plane strain conditions, a computer program was developed in order to simulate
the cyclic undrained behavior of sand. Figure 9 shows an illustrative example of results for
the cyclic triaxial test calculated by the model. The stress ratio in this figure is a quantity
divided by R(S) in Eq.(l5). A gradual increase in pore water pressure can be seen not
only in loading and reloading conditions but also during the unloading process. when the
stress level reaches the phase transformation lines, the effective stress path goes up along
the failure lines, which is one of the typical features of liquefied sands.
0.6~-------------------------,
- U
Q.
~ 0.4
.E:: 0.2
0'
o o.o~--~~~~++H+~~u-----~
....
~ -0.2
0::
Ul -0.4
Ul
(j)
t -0.6 0
100 200 300 400
til
Effective Confining Pressure (kPa)
Figure 9. Effective stress path of an element

The equations governing the interaction of porous solid and liquid were first established
by Biot in 1941 and then the equations were extended to dynamic problems. Zienkiewicz
and Shiomi (1984) applied these equations to nonlinear finite element analyses. The

gO[V~~ing~q](u?~}n~Of[~:~i~' ~~~, S:l~~]d ~:}e :u[~ ca~ ~e:n} t:n{~ns}ilie(::~;


M f Uf _ C +K
2 2
C +K
,,2 3
UfO o_flU f ff
where M ; mass matrix, C and K ; viscous matrices
K ; Stiffness matrix f; external forces
U; absolute displacement (vertical and horizontal component).
Subscripts sand f indicate the components for solid and liquid, respectively.
Pore pressure can be introduced from compatibility equations and defined as
P=Pfgll{ (1-n)U si,i+nU fi,il (19)
where n ; porosity g; the gravity acceleration Pf; the density of water. 11 is a parameter
introduced to approximate the incompressibility of fluid and, in the present analysis, it is
assumed to be 1011. Direct integration of Eq.(18) was made by Newmark's method
assuming a=O.3025 and b=O.6.
A one dimensional response of a soft layered foundation was calculated on the site
where pore pressure and seismic ground motion had been observed. Although there exists
tuff rock at the depth of 58m, where the velocity record H2 was obtained, the gravel layer
at the depth of 48m was selected as the base layer for input waves since it has a shear wave
velocity of 700m and is rigid enough compared with the surface layers. The soft surface
layers were divided into 26 elements and the thickness of each layer was held thinner than
3m in order to raise the accuracy of the calculation. The observation records of the
464

Fukusimaken-Oki Earthquake on February 6, 1987 were chosen as the object of the


analysis.
Horizontal component of H2 motion perpendicular to the direction of the epicenter was
used to reproduce the input wave. The velocity response of the ground surface (HI) to the
input wave was shown in Figure 10, where the fine line indicates the calculated result and
the bold line shows the observed one. Figure 11 indicates the time histories of pore
pressure at PI, P2 and P3 points, respectively. In this figure again the calculated results
were plotted by fine lines and the observed ones by bold lines. It should be noted that pore
pressure records obtained by piezometers have vibrational components and only excess
pore pressure was extracted by using a moving window of 4 seconds in the time domain.
Although slight differences can be recognized in the maximum values and patterns of
generation and dissipation of pore pressure, good agreement was obtained by the
simulation.
u
Q)
<II 5 -----~--------------,

---
E
U
Hl

>,
+-' 0
-rl
U
a
~ -5 ~, I I

> 0 5 10 15 20
Time (sec)
Figure 10. Velocity response
2
P1 (5. 7m)

~
~

0
-Observed
-1
-Calculated
~ -2
~
2
Q)
P2 (13. 2m)
1
>-0
:l
UJ 0
~
UJ -Observed
~ -1 - Calculated
p.
(lJ
-2
2
"
0
p.

0
-Observed
-1 -Calculated
-2
0 5 10 15 20
Time (sec)
Figure 11. Pore pressure responses

By applying the above equations into a two dimensional FEM model, nonlinear
effective stress response analysis can be made for a cross section of a darn. The two
dimensional analysis was made in the case of a centrifuge shaking test model. Figure 12
465

indicates calculation results of the residual deformation due to plastic deformation after 18
seconds from the initiation of vibration. Figure 13 shows also the calculated pore pressure
response in the embankment model. The residual deformations are crucial to the damage
potential and stability of the dam since it may results in the loss of freeboard. However, the
efficiency and the accuracy of the analysis depend on the ways of modelling and
constitutive relations used in the analysis. Therefore some kinds of calibration will be
needed to evaluate these calculation results by use of field observation results or centrifuge
test data. It is necessary to accumulate the case histories of large permanent deformation in
dams which is caused by actual earthquakes.

Scale of Embankment Scale of Displacement


12
(mm)
0.12 r I

~mm)
o 160
~~~~~~~~~~~~ 0 . ~5

Figure 12. Residual deformation in a dam model


L 1m. 1 01
('0

30

o
cc
~ -30

Elm.45

30 60 90 120 150 180


Time (mSec)

~=;o IS=:J~
Elm.4S Elm.71
Figure 13. Pore pressure response in a dam model
466

6. Earthquake Resistant Design of Fill Dams in Japan.


6.1 FACfOR OF SAFETY

The safety of rockfill dams to seismic forces is usually examined by conventional static
slip circle method. In this method seismic forces are considered as static body forces
which act on the dam body in horizontal and vertical direction respectively. These seismic
body forces are expressed by a ratio the gravity force which defines two seismic
coefficients of the dam, namely the horizontal seismic coefficient kh and the vertical one
kv. Although the seismic coefficient is sometimes related to the maximum acceleration of
the earthquake, there are few well established relations which can explain the destructive
effects of the seismic forces on the structures. In the design criteria established by the
Japanese National Committee on Large Dams, the values are classified by types of dams
and zones in which the dams are to be constructed. The design seismic coefficient
applicable to the dam depends on the dam type as listed in Figure 14.
Stability analysis is made by the pseudo-static method in which the factor of safety is
calculated based on conventional circular slip surface method. The stability of slopes must
be checked for five cases in which reservoir water is; I) at high water level, 2) between
high and low water level, 3) at surcharge level, 4) in rapid draw down and 5) reservoir
empty. One half of seismic coefficient may be taken when the reservoir is empty and then
the water is at surcharge level. The safety factor is calculated from effective stress on the
slip surface, therefore consolidated undrained test results were used for the decision of the
design strength parameters of embankment materials.
fZ:LJ Zone I
o Zone II
CJ Zone III
Zone' Zone" Zone "'
Uniform
Type 0.15 0.15 0.12
Zone
0.15 0.12 0.10
Type

0/
Figure 14. Zonation of seismic coefficient

6.2 DYNAMIC ANALYSES

The safety of a dam shall be confirmed by means of dynamic analyses. For the input waves
used in the response analyses, the strongest earthquake predicted as shaking the dam is
taken. The maximum acceleration and predominant period of the earthquake are estimated
from the magnitudes and epicentral distances of historical earthquakes and foundation
conditions. Empirical equations for the maximum acceleration observed at rock foundation
are often used to evaluate those of input waves. An empirical equation proposed by
Okamoto is frequently used in Japan.
am (.1 + 40) 2
log I a 6 40 = 1 0 0 ( - 7 . 6 0 4 + 1. 7 244M - O. 1 03 6M )
467

where urn is the maximum acceleration, 11 is epicentral distance and M is the magnitude of
the earthquake source. Two or three waves having characteristics similar to the earthquakes
observed at the dam site are chosen and modified to fit the maximum value.
An artificial earthquake wave may also be used for a input wave. Dynamic analysis is
usually made for a two dimensional finite element model of the maximum cross section of
the dam. Non linear analysis or equivalent linear analysis shall be executed to predict stress
state in embankment due to earthquake loading. After adding the dynamic stress to static
one, local safety factor FSL will be calculate for each element. The local safety factor is
defined by the equation
2C cos <)l + (cr , + cr 3 - 2 U) sin <)l
F~ = - - - - -cr,-cr
- - 3- - - - -

o
Figure 15. Definition of local safety factor

which is derived from the stress state shown in Figure 15. If there appears many
elements, whose safety factor is less than unity, as a block distributed in a region, and
continued for a while, slip line going through the mobilized plane of the elements must be
checked. If any continuous potential slip surface might be drawn by connecting the lines of
potential slip lines of mobilized plane of each element, then the factor of safety along this
slip surface will calculated.

7. Conclusions
It could be said from the experience of major earthquakes that well designed rockfill dams
are obviously stable even if they are strongly shaken by earthquakes, and engineers can be
encouraged by the good performance of well constructed rolled fill dams. On the other
hand there are a lot of cases of failures of dams, the causes of which can not be
investigated by any means. In order to ensure the seismic stability of rockfill dams, the
following studies must be executed:
1. To investigate the dynamic behavior of embankment, observation of earthquakes on the
existing dams should be strongly recommended in the seismic regions.
2. Three dimensional response analysis will be necessary to estimate the deformation in the
transverse direction, longitudinal direction and up-down direction as well.
3. For the purpose of 3-dimensional analysis comprehensive constitutive relation should be
used, by which 3-dimensional shear behaviors as well as time dependent behaviors can
be taken into account.
4. Dynamic behaviors as well as static deformation characteristics of fill materials should
be exactly investigated by means of large scale laboratory testing apparatus.
468

5. Based on solid-liquid interaction analysis, dynamic nonlinear behaviors of embankment


must be investigated, which can predict the magnitude of residual deformation which
may cause the loss of freeboard.
6. Engineers must keep in the mind that the stability of the structures are always assured
not only by good design but also by careful construction of embankment

Ad;nowledgement

The author wishes to express his sincere appreciation to Professor M. Kamiyama of


Tohoku Institute of Technology and Dr. T. Shimizu of Kumagaigumi Co. Ltd. for their
kind assistance in the observation. He also wishes to extend his gratitude to Dr.Baba of
K.Baba Engineering Consultants Co.Ltd. for his kind advices on this study. This study is
partly subsidized by the Grant -in-aid of the Ministry of Education of the Japanese
Government. The financial support is gratefully appreciated.

References

Akiba M. and Senba H. (1960) : The earthquake and its influence on reservoirs
in Akita prefecture, J. Agric. Engg. Soc. Japan, Vo1.13 No.3 pp.31-59 (1941)
Ambraseys N.N. (1987) : On the seismic Behavior of Earth Dams, Proc. of 2nd
WCEE,VoLl pp.331-358
Arulanandan K., Yogachandran c., Muraleetharan K.K., Kutter B.L. and Chang
G.S. : Seismic Induced Row Slide on Centrifuge, Proc. ASCE, Vol.114 No.12,
pp.1442-1449
Baba, K. and Watanabe, H. (1979) : On A consideratio for An Earthquake-resistant Design
Method for Rockfill Dams, Proc. of 13th Congo on Large Dams, New Delhi, Q51 R15
pp.l049-1074
Casagrande,A. and Shanon : Strength of soil under Dynamic Loads, PrOC. of
A.S.C.E. Journ. of S.M.F.E., Vol.74 No.4, pp591-609 (1948)
Chopra A.K.(1967) : Earthquake Response of Earth Dams, Proc. of ASCE, Vol.93
No.SM2, pp.65-81
Clough and A.K. Chopra (1963): Earthquake Stress Analysis in Earth Dams,
Proc. of ASCE Vol.2 No.EM2 pp.197-211
Dafalias,Y.F. and Popov, E.P.(1976) : Plastic internal Variables Formalism of Cyclic
Plasticity, J. of Applied Mechanics, pp.645-651
De Alba P.A., Seed H.B., Retamal E. and Seed R.B. (1988) : Analysis of Dam
Failures in 1985 Chilean Earthquake, Proc. ASCE, Vo1.114 No.12, pp.1414-1435
Finn W.D.L. (1989) : Permanent Deformations in Ground and Earth Structures
During Earthquakes, Proc. 9th WCEE, Vol.8 pp.201-212
Finn W.D.L., M. Yogendrakmar, R.C. Lo and N. Yoshida (1989) : Direct
Computation of Permanent Seismic Deformation, Proc. 9th WCEE, Vol.8
pp.219-224
Gazetas G. (1987) 'Seismic response of earth dams: some recent developments',
Soil Dynamics and Earthquake Engineering, Vol.6 No.1, pp. 2-37.
Ghaboussi,J. and Momen H. (1982) : Modelling and Analysis of Cyclic Behavior of
Sands, Soil Mechanics - Transient and Cyclic Loads, pp.313-342
Gillon M.D. and Newton C.J. (1989) : Earthquake Effects at the Matahina Dam,
New Zealand, Proc. of Discussion Session on Influence of Local Soil
Conditions on Seismic Response, 12th ISSMFE, pp.37-46
469

Goodman R.E. and Seed H.B. (1966) : Earthquake-induced Displacement in Sand


Embankment, Proc. Of ASCE, Vol.92 No.SM2, pp.125-146
IIda, K (1938) : The Velocity of Elastic Waves in Sand, Bull. Earthquake Research Inst.,
Vo1.16 pp.131-144
Kamiyama M. and Yanagisawa E. (1986): Model for Estimating Response
Spectra of Strong Ground Motions with Emphasis on Local soil Conditions,
Soil and Foundations Vo1.20 No.2, pp.16-32
Kawakami F., Asada A. and Yanagisawa E. (1967) : Behavior of an Earth Dam
during Earthquakes, Trans. of 6th Int. Conf. on Large Dams, Vol.4 pp.19-37
Khalid S., Singh B. Nayak G.c. and Jain O.P. (1990) : Nonlinear Analysis of
Concrete Face Rockfill Dam, Proc. ASCE, Vol.116 No.GE5, pp.822-837
Kutter B.L. (1982) : Deformation of Centrifuge Models of Clay Embankment Due
to Bumpy Road' Earthquakes, Proc. of the Conf. on Soil dynamics and
Earthquake Engineering, Vol.l pp.3 31-349
Kutter B.L. (1984) : Earthquake Deformation of Centrifuge Model Banks, Proc.
ASCE, Vo1.110 No.12, pp.1697-1713
Logani K.L. (1981) : Design Measures to Improve Performance ofFill Dams
Under Earthquake Loading, Proc. of Int. Conf. on Recent Advances in
Geotech. Earthquake Engg. and Soil Dynamics, Vol. 1 pp.465-471
Matsumoto N., Yasuda N.,.Ohkubo M and Arakawa S. (1987) : Dynamic Shear
Strength of Coarse Grained Granular Materials, Proc. of Int. Symp. on
Earthquake and Dams,ICOLD
Mononobe N., A.Takata and M.Matunuma (1936) : Seismic Stability of The
Earth Dam, Trans. of 2nd Congress on Large Dam Vol.4
Mroz,Z. and Norris, V.A. (1982) : Elastoplastic and Viscoplastic Constitutive Model for
Soils with Application to Cyclic Loading, Soil Mechanics - Transient and Cyclic Loads,
pp.173-217
Newmark N.M. (1965) : Effects of Earthquakes on Dams and embankments,
Geotechniq ue V01.15 No.2, pp.139-160
Paskalov T.A. (1985) : Earthquake induced deformation on Earth-fill and Rock-
fill dams, Soil Dynamics and earthquake Engineering VolA No.1, pp.35-42
Penman A.D.M. (1986) : On the Embankment Dam, Geotechnique, Vo1.36 No.3,
pp.303-348
Poorooshasb, H.D.and Pietrusczak, S. (1985) : On Yielding and Flow of Sand, A
Generalized Two-surface Model, Compo Geotech. VoLl, pp.33-58
Rocha M. (1957) : The Possibility of Solving Soil Mechanics Problems by the
use of Models, Proc. 4th ISSMFE, VoLl, pp.183-188
Sarma S.K and M.R. Barhosa (1985) : Seismic Stability Analysis for Rockfill
Clay Core,Geotechnique, Vo1.35 No.GE3, pp.319-328
Seed, H.B. and Chang, C.K (1966) : Clay Strength under Earthquake Loading
Conditions, Proc. of ASCE, Vol.62 No.SM2, pp.53-78
Seed H.B. and R.W. Clough (1963) : Earthquake Resistance of Sloping Core
Dams, Proc. of ASCE Vol.89 No.SM1, pp.209-242
Seed, H.B., Lee, KL., Idriss, I.M. and Makdisi, F.I. (1975) : The Slide in The San
Fernando Dams during The Earthquake of February 9, 1971. J. of GED.,ASCE.
Vo1.101 No.GT7, pp.651-688
Seed, H.B. (1979) : Considerations in the Earthquake Resistant Design of Earth and
Rockfill Dams, Geotechnique Vol.29 No,3, pp.213-263
470

Seed R.B., Dickenson S.E., Riemer M.F., Bray J.D.,Sitar N., Mitchel J. K., Idriss
LM., Kayen R.E., Kropp A., Harder L.F. and Power M.S., (1990) ;'Preliminary
Report on The Principal Geotechnical Aspects of The October 17, 1989 Lorna
Prieta Earthquake', EERC Report No.UCB!/EERC-90!/05
Shibata T.,Kawakami F. ,Yanagisawa E. and Katayama T. (1971) : Damage to Soil
Structures, General Report on the Tokachi-Oki Earthquake of 1968, pp.675-720
Shimming B.B., Heas H.J. and Saxe H.C. (1966) : Study of Dynamic and Static
Failure Envelopes. Proc. of A.S.C.E., Vo1.92 No.SM2, pp.105-123
Tatsuoka, f., Iwasaki, T and Takagi, Y (1978) : Hysteretic Damping of Sands and Its
Relation to Shear Modulus, Soils and Foundations, Vol.95 No.2, pp.25-40
Taylor, D.W. and R.V. Whitman (1947) : The Behavior of Soils under Dynamic
Loading, M.LT. Report to Chief of Engg. U.S. Army
Troncoso J.H. (1989) : The Chilean Earthquake of March 3, 1985: Effect on Soil
Structures, Proc. of Discussion Session on Influence of Local Soil Conditions
on Seismic Response, 12th ISSMFE, pp.l-lO
Wong,R.T., SeedH.B. and Chan C.K. (1975): Cyclic Loading Liquefaction of Gravelly
Soils, Proc. of ASCE.,J. GTD. VoUOI No.GT6,
Whitman R.V. (1966) : The Behavior of Soils under Transient Loading, Proc. of
4th Int. Conf. on S.M.F.E., VoU pp.207-212
Woods R.D. (1963) : Preliminary Dosion of Dynamic Static Shear Apparatus
for Soils, Technical Documentary Report, No.RID IDR 63-305
Yanagisawa E. (1982) : Effect of Ground Condition on Vibrational
Characteristics of Earth Structures, Proc. Of JSCE No.324, pp.101-110 (in
Japanese)
Yanagisawa E. and Fukui T. (1980) : Performance of the Tarumizu Rockfill Dam
during Earthquakes, Proc. of 7th WCEE, VolA pp.133-140
Yanagisawa E., Lee W.S. and Ohmura Y. (1984) : Seismic Stability Analysis of
an Embankment, Proc. of 8th WCEE Vol.3 ppA13-420
Yanagisawa E., Ohmiya H. and Shimizu T. (1987) : Seismic Response of Pore
Water Pressure In surface Sand Layer, Soil Dynamics and Liquefaction,
Elsevier, pp.221-229
Yanagisawa E. and Sugano T. (1989) : Measurement of Seismic Induced Pore
Pressure, Proc. of 12th Int. Conf. on SMFE, Vol.3 pp.2023-2026
Yziquel A., Lino M., Post G. and Tardieu B. (1981) : Seismic Analysis as a Tool in
the Design of Two Earth Dams, Proc. of Int. Conf. on Geotech. Earthquake Engg. and
Soil Dynamics, VoU pp.395-400
Zelikson A., Devaure B. and Badel D. (1981) : Scale Modelling of Soil Structure
Interaction during Earthquakes Using a Programmed Series of Explosions
during Centrifugation, Proc. of Int. Conf. on Recent Advances in Geotech.
Earthquake Engg. and Soil Dynamics, Vol. 1 pp.361-366
CH APTER 16
MONITORING AND SAFETY EVALUATION OF ROCK FILL DAMS
A. VEIGA PINTO

I. Introduction

The purpose of a safety evaluation is to determine the status


of a dam in relation to its structural, operational and
environmental conditions . The evaluation should confirm the
existence of adequate safety conditions or, if problems exist,
detect a nd identify their causes, develop analysis or tests to
study their incidence in dam safety and, if necessary to
recommend operational r estricti ons, remedial repairs or both.
The act ual performance of the dam should be analysed taking
into account the dam and appurte nant wo rks. This analysis may
lead to a revision of the design criteria or hypothesis if the
actual behaviour disagrees with the intended one .
usually the engineer prepares the dam design by using deter -
ministic methods, i . e . , he cons iders that the probability of
failure of the dam is to be the minimum possible (10'4 for
instance) since failure o! such a structure has results that
may often be catastroph ic. Nevertheless numerous fai lures have
occurred throughout the history of dam construction, which has
prompted a reflection on the mistakes behind them, and thus
has contributed to improving the tec hnique s and methods of dam
design and constructio :-, .
In dam engineering, the assessment o f safety is therefore
one of the most important tasks. safety means to assessing the
level of risk, which is considered as the probability of
occur rence of a deterioratio n . This probability depends on
care taken and knowledge duly applied in design and
construction. For instance, in the structural design the
engineer defines the most economical prototype with a very low
probability of occurrence of an incident.
The analysis of deteriorations in dams resulting from their
abnormal performance s hows that such p erformances are usually
produced by a complex of causes and mech a n isms calling for
great care in their interpretation . This i nterpretation should

'"
E.. Mor(m/,o dOl N~\'rl (cd.). AdwlrIcrs ;f! RaclJill Strlfe/II'ts. 471-522.
e 1991 KIIII''f',A cod~mic PllblishuJ.
472

be based on the measurements of the instrumentation and of


visual inspections, but it does not preclude an adequate
engineering judgement based on experience.
The monitoring system is important for to analysis of the
safety of the work during its effective life and helping to
develop new knowledge about the structural behaviour of these
works, with a view to the safety and economy of dams to be
designed and constructed in the future.
Diagram of Fig. 1 shows the various steps in assessing the
safety of a dam.

EXPERIENCE

BASED ON SIMILAR
SUCCESSFUL OR
UNSUCCESSFUL
DAMS PERFORMANCE

PLANNING SYSTEMS

THEORETICAL ANALYSIS
(FEM)

Figure 1. Programme for safety evaluation of a dam


First comes past experience, based on the successful or
unsuccessful performance of similar dams, because this enables
us to plan monitoring and provide performance models as well
as standards and rules that can be taken into account. The
instrumentation gives us results of field measurements, such
as leakage, deformation, pore-pressures, etc. Other known
473

physical phenomena such as cracks, fissures, local leakages


are obtained from visual inspection. The actual behaviour can
thus be pictured and compared with the results of theoretical
analysis obtained, for example, with Finite Element Method
(FEM) and using the mechanical properties of the materials.
Analysis of the safety of a dam will be performed, for
example, based on anomalies or disagreement between observed
and calculated values. Often, after picturing the behaviour of
a dam, it may be necessary to re-analyse certain points, such
as measurements of displacements, or provide new theoretical
stress-strain prediction.
Dam safety is of fundamental importance, and can be ensured
when the behaviour of the dam is fully understood and
constantly monitored by suitable instrumentation.

2. Type of Measurements

To analyse the safety of a dam it is necessary to choose the


physical and mechanical quantities to be measured. The concern
put in safety analysis depends mainly on the importance of the
dam, which is related to its height, and on the risk factors
associated with it. As regards this aspect, the International
Commission on Large Dams (ICOLD) in 1982 published its
Bulletin 41 (ICOLD, 1982) in which a proposal is presented for
assessing the risk condition of a dam, as shown in Table 1.
The Portuguese members of the ICOLD Task Group ("Automated
observation for the safety control of dams") working to the
Committee on "Deterioration of Dams and Reservoirs" suggested
some modifications to the prescriptions of that Bulletin
(Silveira et al., 1983) which are presented in this paper.
The risk factors associated with a dam are related with
external or environmental conditions, which by their nature
are difficult to forecast, with dam condition/reliability, and
with human/economic potential hazard.
The factors due to external or environmental conditions (E)
are connected with: the seismicity of the region, the
possibili ty of embankment slidings into the reservoir, the
risk of floods higher than the design floods, the reservoir
function (type of storage, management) and the occurrence of
aggressive environment actions (climate, water). The factors
due to reliability (F) depend on: the characteristics of the
design, quality of the foundations, reliability of flood
outlet equipment and maintenance conditions. As regards human
and economic factors (R) they are related with the retention
capacity of the reservoir and the type of human settlement
downstream of the dam.
From the mentioned Table the factors E, F and R may be
defined for each dam, corresponding to the arithmetical
average of the partial indices. The overall risk factor is the
product of the values of those three indices.
The physical quantities to be measured are mainly related
wi th the risk factors. One of the most important is the
.j:>.
-.J
.j:>.

TABLE 1. Proposed risk conditions evaluation

Human/economic potential
External, or environmental Dam condi tionjRel iabi 1 i ty (factor F)
COnd~tiOnS ~=:_ E} _______ _+__ hazard (factor R)
.----
. Danger of Danger of I Reserv. func. I Aggressive I
F d t' I Flood outlet Malntenance Reservoir
t
I Downstream
P~~~~~l Seismici ty b:~~e::~~~:s hf~~~:Sthan (;~~:, O!a~~~~J a~~~!~~n7~~~_1 pr~~~~~1~~~~g oun a lons
I
equipment condi tions vo~u~~af:J) installations
(0:,) design floods ment) : mate, water) I

f------ (1) (2 ) (3 ) (4 )
. 1:;-; ---. -1---i0----t-0;-i7~) 19) (l0)

r Very low I Pluriannual,


~
Uninhabited
Minimal Minimal I robabilit annual or : , . < lOS zone of no
or nil or nil [I 'icancre~:m:)~ ~~~~~~:' 'Very weak I Adequate Very good Rehable Very good I economic
(v < 4 cms-') I importance
-- - -------- ------

Low Isolated
< v < 8
ems-') Low . _ _ ___ +~~_~_ - I
Small towns,
Middle Very low I . ' 7 agr ieul ture,
I probability Weekly
r------
Middlefcceptable Acceptable
'"~ :i~W' -:=;::;,.'"
Satlsfac-! 10" - 10
I
I craftsmen's
< v < 16, I, I 1fill damS): storage tory I acthi ties
--- ems
-) I, !--' ~-- I
st I Middle
rong I Daily Strong , 10 7 - 10 9 I sized towns,
< v < 32 . storage I I small
_ ~I I ___ I ~~~~ries
~ms 1) t
f____ ___
Very , , B i g towns,
strong Pumped i Very Poor I > 10 9 industries, I I
Iv > )2 storage 1 strong I I nuclear
ems') _ _ _ + 1____ _________ _ __________ I installations

6
la)
Danger of
blg sl.tdes
High
probability
I Poor or
bad
Insufficient. Unsatisfae-
Not operatio-
nal
tory
t
(a) Abnorma 1 conditions; technically unavoidable intervations

5
E = I
5 i=1
",
",
11
R = L 0-, Global index: o-g - E.F.R
2 i=10
475

possible occurrence of an uncontrolled seepage and so it is


decisive to measure the seepage discharge in downstream zones.
Also the pore pressures in the clay cores should be known,
since a quick increase may be related with important flows due
to fissures existing in the core fill. As is known, this
phenomena may lead to piping mechanisms and consequently cause
harm or can put in danger the stability of the fill.
Another question is structural safety analysis in terms of
stress-strain performance. For instance, measuring stress can
also be very useful, mainly in clay core dams, and when the
material is less stiff than the shell materials, it leads to
important stress transfer and leads to specific problems of
hydraulic fracturing.
For rockfill dams, displacement measurements seem to give
good information about the structural behaviour of the work,
namely, in rockfill dams with upstream facing. Also in zoned
rockfill dams it is important to avoid important discon-
tinuities in the displacements inside the fill. Usually, the
overall behaviour of the dam as regards its displacements can
be obtained by measuring the surface displacements.
Not so important as regards the safety of rockfill dams but
rather as an aid in order to understanding its performance are
the dynamic effects and action such as the accelerations and
displacements due to cyclic loadings. The obtaining of these
parameters are specially important in seismic zones.
Besides the quantities measured by instruments referred to
later on, other evidence should be analysed by visual
Lnspection in qualitative terms, such as fissures, local
dlidings, leakages, etc., which can give indications of
extreme importance for the structural safety.

3. Monitoring Scheme Design

3.1. SELECTION OF MONITORING EQUIPMENT

To obtain proper conclusions about the safety and the beha-


viour of rockfill dams it is essential carefully to prepare a
plan of monitoring activities. This attention must be paid in
order to ensure that the required information will be obtained
during the construction phase and during the life of the
structure. The requirements of the system and the procedures
used for analysing observation data should be formulated in
detail, and selection of the measuring devices and their
location should meet those requirements.
All the dams are different, and therefore, it is not
possible to establish fixed rules for monitoring planning.
Nevertheless, concerning the importance of the dam (related
with its height) and the risk factors, some proposals can be
pointed out as recommended by ICOLD in Bulletin 41. As it was
said above some corrections to the initial proposal are
476

presented in Table 2 as regards the equipment that must be


installed in fill dams (Silveira et al., 1983).

TABLE 2. Proposed monitoring installations and


surveillance scheme

t
Displacements Water pressure
Dam height 1----,-------1 Total II Seepage I -____-,--___ ~ Meteorology Seismology
em) I stress d.lscharge I open electric

.
Surface 1 Internal I piezometer piezometer
-- :-- -T---- -- - ---
J
---

1
< 15 if 0: 9 >15 : Tota~ dl.SC~arge if 0: >10 if Cl: g > 9
or R > 3 i ~r ~g?> 3 or R9? 3
t-----+---- -- t-- ------
and~~

15 to ]0 if 0:
:d :
*>10 if; >20 I! Total discharge if (11=-5

(+-I
or 3 9

-----+--- -- ----+-- ---- -------- ---._-----


,
I I
30 to 50 I If 'erg > 1 0 . ) Part.lal
dlscharge ....
i orR >3
j--___
lfR~3
_ _______ _
1------ _ _ ~~ ___ J . _ _

I
I !

I .----------- ----+- ----~


50 to 100 if R?3
Partial discharge

1------- ---

!
> 100
i Partial discharge
. i i
* Installation must be present
(*) Installation is optional

The recommendation should be accepted in general terms,


i.e., with a certain flexibility, since each dam is a special
case and its own characteristics should influence the
preparation of the observation plan. For example, in the case
of a rockfill dam with central core, the observation should be
concerned with determination of clay core pore pressures
complemented with measurements of stresses in order to assess
the stress transfers between the core and the shells.
Moreover, in the case of rockfill dams with upstream facing
it is mainly the displacements that need to be measured, as
well as other features associated with the facing, such as
opening of joints and extensions of the slabs in the case of
reinforced concrete facing.
One limiting factor that has to be taken into account in
moni toring planning is that measurements are discontinuous
both in space and in time. Whereas the time interval can be
shortened for significant parameters (automation is an
example), it is impossible to provide a very large number of
measurement points around a sensitive zone. It is inevitable
that sensitive zones in places that have not been foreseen
have an insufficient number of instruments (or even none) in
spite of the availability of complex rheological and
mathematical models for investigating dam behaviour.
Discontinuities in the instrumentation records is therefore
477

a very actual and important failing. Its seriousness depends


on the type of instrument and the distance between them. It is
necessary to use the concept of confidence interval of the
Probability Theory to quantify the risk involved. Risk is also
influenced by quality control of the materials of the dam and
foundation during construction.
In monitoring planning the designer may be involved since he
knows what kind of information is required for safety
analysis. He may point out the critical zones and mechanisms
that may endanger the safety of the work. The instrumentation
engineer will recommend what instruments and operations can be
used in order to get the best answers as regards the behaviour
of the dam. Some points should also be included about the
contractor's construction methods.
When planning monitoring, it is convenient to decide that
during the first filling the rise of the water level shall be
stopped at predetermined levels in order that a complete set
of measurements can be made. However, this does not always
happen or only happens partially as in the case of dams in
arid zones built for flood regulation whose reservoirs are
filled only by river floods of relatively short duration.
As an example, Fig. 2 shows the devices installed in the
maximum cross-section of Beliche Dam, a 54m high central clay
core rockfill dam (Veiga Pinto, 1983a). The instrumentation
of the dam is: total-stress cells, intended for measuring the
transfer of stresses between the core and the shells,
hydraulic piezometers installed in the central core,
inclinometers for measuring vertical and horizontal
displacements inside the dam, and surface survey monuments to
record displacements at points on the crest and faces.
Another example lS presenLed in Fig. 3, regarding Apartadura
Dam (Veiga Pinto and Silva Gomes, 1988). This dam now under
construction will be 46m high and will be impervious with an
asphaltic membrane on its upstream face.

3.2. SELECTION OF INSTRUMENTS LOCATIONS

The selection of instrument locations should take into account


the predicted behaviour, especially regarding critical zones.
Locations shuuld be selected in such a way that data can be
obtained as early as possible during the construction process.
Flexibility should be maintained so that planned locations
can be changed as new information becomes available during
construction. A very recent case in a Portuguese dam serves to
ilJ.ustrate this.
One of the first Portuguese developments concerning a
complete type A predictions was made for Beliche Dam (Veiga
Pinto, 1983; Veiga Pinto and Maranha das Neves, 1985). In the
monitoring scheme all the devices inside the dam were
concentrated in the maximum cross-section, taken as fully
representative of the dam behaviour; the aim was to compare
478

t Inclinometers
o 30m
~I=!;=;;;;;l=~1

Cluster of pressure cells


V Casagrande type piezometers
, Surface survey monuments

Figure 2. Instrumentation installed in the maximum cross-


-section of Beliche Dam

598.00
~
585.00
.~
576.00
-~

o 10 m
1====*

~ _Inclinometers

~_5ettlement and horizontal displacements device

.. _ Surface survey monument

Figure 3. Instrumentation installed in the maximum cross-


-section of Apartadura Dam
479

the calculated quantities with the respective monitoring data


and to provide cross-checks among instruments. Nevertheless,
at the end step of the first filling an exceptional flood for
a return period of 1000 years occurred. This fact, coupled
with the lack of a concordance between predicted and
registered settlements, led the staff responsible to improve
the piezometric monitoring network by increasing the number of
hydraulic piezometers from 8 to 46 to be installed in a
further 8 cross-sections in order to cover the different zones
of the clay core dam (Veiga Pinto et al., 1988).

3.3. INSTALLATION PLANS AND PROCEDURES SUBSEQUENT TO CONS-


TRUCTION PHASE

As has been said, good results in a dam monitoring depend on


a well balanced combination of equipment and skill of the
people to install it. The importance of an efficient job in
installing the equipment by responsible technicians in order
to obtain reliable values for the behaviour of the structure
has been emphasized and, therefore, installation procedures
should be written in advance of scheduled installation dates.
These installation procedures should include a detailed
listing of materials, tools, machinery and manpower required
to execute the work.
Plans should be coordinated with contractors and
arrangements provided to protect installed instruments from
damage caused by construction activities.

3.4. MONITORING FREQUENCIES

Bulletin 41 of ICOLD also gives the frequency of measurings


that should be carried out in the instruments installed in
fill dams. This frequency is presented in Table 3 including
some suggestions of the Portuguese members (Silveira et al.,
1983).
The proposed values referred to above must be adapted to
each particular case. In rockfill dams with upstream
impervious facing, for instance, some of the quantities of
Table 3 are not very important to be measured, namely the
water pressure and total stresses and consequently they can be
discarded. On the other hand when uncommon conditions occur,
such as abnormal evidence not predicted, very rapid or very
slow fillings and earthquakes, the relevant adaptions should
be adopted.

3.5. PLAN OF FIRST FILLING


As the first filling of the reservoir corresponds to the first
loading test of the dam, when the reservoir levels reaches the
maximum, the product of the costs (economic or human) of a
.jC.
OG
TABLE 3. Proposed frequency of readings for fill dams C

Se.epl!!t.qa
DlsplaeOlPont.. Water preSBu.["e Visual inspections
H'oi.9bt Lite ~-----r----~I Total ~__d_l._e, h_. _r9_'__4-________~______-4______-r______r-______~ Rosorvolr
:$t.aqe stresses level Met e orOl09Y
(.)
Surface I Intern'lll open electriC Noraal I Spoeial ! Xcepcioni!t.l
Tot,a l IPorthll pis:i!oac.tor
IplezollLOter
< IS
< I-- d_ -
.j I-
15 to :10 _~-~ I :-R- ~ 4 I d
u
10 to SO
I I : i=d d

50 to 100
.t --+- f----+-
8 I -W
. I- w --+ ~ ~ --+- j - - -I- ~ --
::> laO Q w w I K d
< 15 5 or Q 5 or Q , or Q
g' ~ d
~ 0
<> A- 6 or 0
--- '- --1
15 to 10 :::: -g or 60r MI60rM 60rH 60rQ 60rM 60r" d
~
I-- . /---II~- ------ I-- ------+-
30 to 50 ~" " 6 or A 6 or K 6 or Q 6 or M ~or A d d
~o: -- -r- --
50 to 100 !: ~ 6 or hllJ. 6 or 1'1 6 or!1IIo 15 or _SA r --+--.__ d
.. ~
0'
:). 100 60rw 60rH 60rw 60r'O

< " B SA SA SA B d d
1l - ---/-- - -- ~-

15 t.o 30 ... -;: A SA ~ _~ ___ 0 -f-~ Q A_ ~ d _ d ___


30 t.o SO ~ 8. II. SA SA H. K H SA K A -4 d d
<:: -. - - - - - - - -f-- - ._--/- -- --- 1-- - ~
SO to 100 r:.!: A Q OM M H '0 H A -4 r d
o w
> 100 ~ SA Q 0 hm h_ hili Q h= SA -4 [" d I
< 15 Q<I S' SA
SA f-
~I ~ d d
SA Q Q SA
J-15 to 30
&
o I o 4 -----+-_~
t d
JO to 50
~ 8 - + SA ~ SA H Q I --"A_ _ I-- K -+-~ d d
5 0 to 100 < - d :
I- ). 100
.~ A
A
~
Q
OHM
o I h. h.. h.
~~
o n. A
_J = d

3 - In the middle and at the end of construction r - Register


4 - After occurrence d - diary A - Annual
5 - At the beginning and at the end of first filling or drawdown w - weekly B - Biennia 1
6 - At the beginning, levels and at the end of first filling or hm - half-monthly QQ - Quinquennial
drawdown
M - Monthly Q - Quaterly SA - Semi-annual
7 - Weekly to several times a week
481

possible rupture by its probability assumes the largest values


during the entire life of the structure.
A first filling plan should therefore be prepared in order
to control the performance of the dam during that phase.
That plan should contain indications about:

1 - procedures to be used in visual inspections;

2 - physical and mechanical quantities to be monitored


in order to assure quick safety control;
3 - frequency of measurements taking into account the
filling programme;

4 - definition of the reservoir levels to be maintained


during a certain period during which a set of
observations should be made to contribute to
assessment of the safety condition of the structure
and to recommend going ahead to the next filling
stage or halting the filling;
5 - safety models to support the safety evaluation.
In some countries it may be very difficult to impose a first
filling plan, for instance in tropical areas where rain is
scarce and sudden large floods may occur. Nevertheless, in
temperate regions, such as Portugal, this philosophy of action
has been acquiring more acceptance and first filling plans
have been implemented in the last few years (Veiga Pinto and
Silva Gomes, 1986; Silva Gomes and Veiga Pinto, 1986).

4. Monitoring Equipment

After defining the type of measurements that can be obtained


in a rockfill dam, a decision must be taken on what type of
equipment and which model's manufacturer to choose.
The number and types of instruments and devices available
have increased and range from simple, inexpensive mechanical
devices to highly complex and costly electronic equipment.
As a general rule, it is sound instrumentation policy to
make the monitoring system as simple as possible, for
obtaining the required data. Unsophisticated but reliable
mechanical pneumatic and hydraulic systems are therefore often
preferred, though sometimes it is advantageous to have a more
sophisticated (usually electrical) measuring system that will
enable remote observations to be made.
Problems of exposure to weather, deterioration in an
aggressive environment, protection against vandalism should
all be considered.
When evaluating a prospective monitoring system, the total
cost has to be considered. This should include not only the
4X2

cost of the instruments but also installation and observation


costS. The reason for this is that in several monitoring
schemes the actual material and installation cost is low, but
the cost of obtaining reading is relatively high, or vice
versa.
Accuracy and sensitivity requirements for monitoring systems
are dictated by geotechnical and structural criteria, relevant
to the project that has to be instrumented. It is useful to
make a rough estimate of the magnitude of the various
parameters to be monitored, by a stress-strain behaviour
prediction using a mathematical model. The instruments should
have the proper measuring range. Accuracy is usually related
to range.
Proper assessment of operational requirements and enviro-
nmental conditions is essential if a monitoring project is to
be successful. Selection of instruments must be based on
important details such as robustness, watertightness,
resistance to corrosion. insensitivity to temperature changes.
long-term stability and type of output signal preferred. As a
general rule, inconvenience to the con~ractor caused by pipes,
casings, cables and other installations must be avoided
wherever possible.
When selecting a monitoring system it is also very important
that the methods of installation and readout facilities
employed should be compatible with the technical
qualifications and skills of the personnel who are in charge
of the installation of the system and carry out the
observations.
The various types of measurements to be obtained in rockfill
dams and the possible candidate instruments are shown below.
4.1. TRIANGULATION AND TRILATERATION NETWORKS

The performance of the external surfaces of an embankment dam


can often provide clues to the behaviour of the interior of
the structure. For this reason, the importance of measuring
the surface displacements of the embankment by geodetical
equipment should be emphasised. Triangulation and trilatera-
tion networks are used for that purpose.
Triangulation and trilateration networks are sets of points
(stations) materialized by concrete pillars (with adequate
foundations, thermically well insulated and equipped with
special forced centering pieces) and by survey targets,
located at regular intervals along the crest and several
berms. In Portugal, it was usual to measure surface
displacements at points in downstream berms, but now. owing to
the increasing importance given to the first filling phase,
displacements at points of the upstream slope are also
measured in Portugal (Fig. 41 (Veiga Pinto et al .. 1988).
Fig. 5 shows the geodetic network of Beliche Dam, which has
two references (PJD and PJE) built outside the construction
4~3

area on bedrock in positions where displacement are not


expected to occur (Casaca, 1987). Fig. 6 shows a reference
pillar that has a Kern centering instrument plate in its upper
base.
The observation network consists of angular and distance
measurements with planned operations and suitable equipment.
LNEC uses ME 3000 Mekometer distanciometers and Kern E2
electronic theodolites. The data obtained are checked and
converted into displacements by appropriate mathematical
models.

~igure 4. upstream face berm of Beliche dam, designed for


installing survey points at 85% of its height.
4.2. PRECISION LEVELLING
Precision levelling consists of determining changes in level
between points fixed to the structure and fixed points with
known displacements.
One precision level (wild Na3 or Wild Na2) fitted with
optical micrometer and invar rods is the equipment used for
measuring vertical displacements of the bench marks. Lately,
however, attempts have been made at LNEC to obtain the
settlements from the vertical angles measured by theodolites.
In precision levelling methods it is possible to obtain a
precision of about a few millimeters.

4.3. INCLINOMETER
Modern inclinometers or slope indicators have been used
4~4

extensively for measuring the variation of inclination along


a pipe in the body of rockfill dams by means of which the
horizontal displacements can be computed.
Since the introduction, in the early 1950's, of the
pendulum-actuated inclinometer (slope indicator) I operating in
grooved plastic casing Ithe same basic concepts have been
applied by several manufacturers and institutes.

&-Fixed survey station

"-Survey station
"-Target

PJ 0 PJE

Figure 5. Geodetic network of Beliche Dam

The four basic components of the inclinometer system are a


movable borehole sensor J a portable digital indicator, an
interconnecting electric cable and a slope indicator guide
casing.
The movable sensor provides an electric signal proportional
to the angle of inclination from its vertical axis. It
contains two servo-accelerometers mounted with their sensitive
axes 90 apart. When operating within the casing, the sensor
is supported laterally by guide wheels running in two
symmetrical grooves (the casing has four grooves, equally
spaced around the inside circumference), and is suspended by
the interconnecting cable.
485

The digital indicator is a portable instrument whose display


shows the angle of inclination from the vertical in terms of
2.5 times the sine of that angle.
A heavy-duty multi-conductor cable with a stranded steel
cable at its center is normally used to provide repeatable
vertical control. Initial sets of readings are taken at
specific depths within the casing (say every 0.5 m) and at
periodical time interval readings are taken at the same
depths. The difference in successive readings at the depths
represents a change in inclination, which is converted to
linear displacement as is shown in Fig. 7. A progressive
change in readings indicates a zone of movement.

Figure 6. Reference pillar with a Kern centering instrument

The cable is attached to the sensor by means of a waterproof


connector.
The casing is a plastic, steel or aluminum pipe, installed
in boreholes drilled in the embankment or fitted during
construction.
The casing installation is usually nearly vertical but in
the upstream facing of rockfill dams the casing is also placed
at the same inclinatio,l as the upstream slope. Accuracy
diminishes in proportion to the inclination of installation.
The deepest inclinometer installations are at about 300m
(e.g. Chicoasen Dam). Accuracy of the inclinometer surveys is
about:: 5mm per 30m of near vertical installation.
Telescoping joints often have to be used every 1.5 to 3m in
order to accommodate the downdrag on inclinometer casing due
to vertical deformation of the embankment and foundation. The
settlements of the telescoping joints are measured by a probe
provided with a system that successively engages the
telescoping joints.

GRADU_ATED
TOTAL DISPLACEMENT (L L sin oe)
----I
E L EC T R IC A L
CAB LE

ACTUAL ALIGNMENT
OF GUIDE CASING
DISPLACEMENT
ILslnoe
I

COUP LIN :::..G_->M'FiF'il1

1,
1
BOREHOLE
- ~

PROBE: "
CONTAINING i ;-, GUI DE
GRAVITY-
i ',' CASING
-SENSING

:1 ~+-,C___A,,-.SI NG GROVE S
,~

i.[. .,.~"
'I
-'\l ' ~. "'0 .'
'"'1'/

Figure 7. Inclinometer principle


-1-X7

In other cases, magnet markers are placed around the


inclinometer tubes as at Keddara dam (Fig. 8). During dam
construction, the access tube can be extended upwards through
the fill, and marker plates containing magnets can be added at
various fill heights. A reading switch sensing unit is used to
detect the magnets and it is lowered into the access tube at
the end of a measuring tube so that the positions and thus the
settlements of the markers can be measured.

Figure 8. Slope indicator case with a magnet marker

The inclinometers placed during construction have the


serious disadvantage of restricting the movement of earth-
-placing and compacting machinery. They are also subject to
uncontrolled movements due to local fill displacement caused
by machines. For example, the inclinometer tubes placed during
the construction of Beliche Dam, with a clay core compacted
above the optimum water content, underwent successive
displacements and at the end of construction it was no longer
possible to insert the sensor in the casing. Afterwards, it
was necessary to put other tubes parallel to the original ones
in order to measure the horizontal displacements.

4.4. FLUID LEVEL SETTLEMENT GAUGE


As has been mentioned, the vertical devices for measuring
settlement have a serious disadvantage because of the
earth-placing and compacting machinery (Fig. 9). In general,
it is better to keep access tubes and connecting leads for
instruments on a plane parallel to the surface of the fill, so
that they can be placed in trenches during construction and
buried permanently.
The gauge consists simply of an overflow unit buried in the
fill and connected by small flexible tubes to a vertical
standpipe in an instrument chamber.

Figure 9. In order to avoid damage to the monitoring equi-


pment, some recommendations are given
In Portugal, we have not used this type of equipment,
al though at the moment the design and construction of a
prototype is being carried out at LNEC.
Some new equipment to measure the displacements inside
embankment dams are now being tested in Sweden (Cederstrom,
1990). The main objective is to fast and easily establish the
actual top of the damcore all along earth-rockfill dams.
Advanced radar examination with an unique mobile multiple-
-antenna radar is used.
4.5. HORIZONTAL DISPLACEMENTS DEVICE
owing to the same reasons that justify the research on a fluid
4X9

level settlement gauge, devices to measure the horizontal


displacements at points parallel to the surface of the fill
are also under research. Normally, the same points to read the
vertical and horizontal displacements inside the embankment
will be chosen.
Very recently, a device used to measure norizontal movements
at several locations within rockfill embankments was designed
and constructed at LNEC. This device was planned to have an
accuracy of about 5mm. In Fig. 10 the working principle of
this system is illustrated (Taco Emilio, 1987 and Silva Gomes
e"t al., 1987).

r ",' '.,

0'"

DET AIL A DETAIL B


~" 1

Figure 10. Internal horizontal displacement device (Toco


Emilio, 1987)
Extensometric wires are anchored to individual plates fixed
inside the dam and horizontally extended towards the
downstream face where a reading station is installed.
The wires are of high strength and they are always kept in
tension by light weights in order to avoid wire entanglement.
A heavier weight is used at the end of the wires during the
reading operations so as to provide a uniform degree of
tension when making a measurement. This value is the
horizontal distance between the reference mark of each wire
490

and the reference mark of the reading station. The temperature


should also be recorded in order to correct length of the wire
for any temperature change.
The permanent target of the reading station is also movable
owing to the displacements of the downstream face of the
embankment. Its location should therefore be established
during the reading operation by triangulation survey, using a
base line between two fixed pillars installed outside the fill
for reference.

4.6. TOTAL PRESSURE CELLS

strains caused by differential movements may seriously alter


the total stresses in the various zones of the dam. This may
be of particular importance in the impervious zone where
information about any reduction in the total stresses imposed
by the overburden is essential.
If pressure from the reservoir water makes the effective
stresses in the impervious zone fall to zero, hydraulic
fracturing may occur and subsequent erosion may lead to
piping. It is therefore a design requirement that all total
stresses in the impervious zone shall be greater, by a
suitable factor of safety, than the maximum pressure that can
be exerted by the reservoir water.
Many different types of earth pressure cells have been used
in attempts to measure total-stress conditions in dam fill. As
a rule, fluid-filled, flat, stiffer-than-soil cells have
produced the best results. This type of hydraulic cell is
preferred in Portuguese fill dams.
Essentially, a hydraulic cell (Fig. 11) is a rectangular or
circular thin jack filled with oil. This jack is connected by
a metallic tube to a diaphragm pressure transducer.
When the enveloping soil pressure acting on the jack is
balanced by the oil pressure imposed a~ the pressure line, the
diaphragm in the pressurized chamber deflects and allows the
passage of a small quantity of oil through the return line.

~ RETURN LINE

SO I L PRESSURE ~ PRESSURE LINE

y 1 1 1 -.r I

""------- FLEXIBLE DIAPHRAGM

OIL/ i I i i I CUSHION

Figure 11. Gl6tz~ (hydraulic) total pressure cell


491

Despite of the difficulties involved in installing this kind


of instrument, good results have been obtained and its use is
recommended. However, in order to avoid the risk of losing
information owing to damage in any of the instruments, the
measuring arrays should be formed of groups provided with
superabundant apparatus. When used near the center line of a
dam, where the major principal stress is expected to act in a
vertical direction, a minimum of three cells should be used.
Fig. 12 shows such cells and is referred to an array placed in
Beliche dam. On the basis of experience in the measurement of
total pressures in this dam, it was decided to install one
more cell at 45 to the vertical plane. Thus we will obtain
two states of stress which, in a plane strain condition, can
be compared.

Fig 12. Detail of installation of total pressure cells


Some procedures have been introduced in this equipment in
order to obtain more reliable measurements. For instance, it
is not intended to place the cells at points dug into fill to
492

avoid fill arching over the pit which will prevent the true
total pressure to reach the cell. Instead, it is preferable to
install the group of cells in a hollow near the surface of the
fill. The cells should be about two diameters apart and the
surface of the fill smoothed to form flat surfaces for the
cells placed horizontally on planes 45 to the horizontal.
Vertical cells can be placed in shallow slots wide enough to
allow excavation material to be backfilled and compacted by
manual hammers to the density of the adjoining fill. When the
cells have been covered, manual pneumatic hammers may be used
to consolidate the fill, but at least three diameters height
of fill should be put over the cells before normal fill
placing continues above them.
As proposed by Dibiagio et al. (1982) in Norway, remote
reading inclinometers attached to total pressure cells may be
used to ensure that compaction does not move the cells too far
from their intended attitude.
It should also be borne in mind that the presence of a cell
affects the stresses in the ground. Taking into account this
fact attempts have been made at LNEC not only to use hydraulic
calibration but also to discover the response of the cells
buried in soil in test chambers as shown in Fig. 13. It was
found that the influence of the soil on cell measurements may
be very important in some cases.

4.7. PIEZOMETER
Piezometers for embankments are used to measure the static
water pressure or head at various elevations in the foundation
and embankment soils. Four basic types are in cornmon use: open
standpipe, closed hydraulic, pneumatic diaphragm and electri-
cal (vibrating wire) diaphragm.
All piezometers consist of tree parts:

a) the intake filter;


b) a pressure measuring device;

c) the connection between the two.

The intake filter must be strong enough to withstand the


maximum total pressure that will act on it, without
appreciable deformation. It must also be narrow enough to give
the required response time to the piezometer, and have
suitable pores in order to allow correct contact with the pore
water of the soil. In partially saturated fills this is a
particularly important aspect.
The simplest type of piezometer most commonly used in
Portugal is the open standpipe of the Casagrande type. This
piezometer has proved successful for many materials and
493

corrosion as a rule is not a problem. It consists of a tip (of


ceramic material with a low bubbling pressure where the
pressure is taken), from which plastic tubes (protected by
galvanized iron, perforated in the tip zone) extend to the
surface of the dam. The internal diameter of the tube (6mro)
enables an adequate response time to be obtained. Measurement
of the water height in the tube is done by an electrical
device consisting of a cable with two conductors: when in
contact with the water this closes a circuit and the
corresponding signal is read on an ohmmeter.

Figure 13. Cell calibration buried in soil

A schematic intake filter of a foundation hydraulic


piezometers is shown in Fig. 14.
Where the depth to the water level is greater than 100m, the
electric cable of the water level probe becomes vulnerable to
breakage due to friction. Although the open standpipe type is
not satisfactory either in very impervious soils because of
time lag, or in partially saturated soils where results are
494

difficult to evaluate, the simplicity, robustness and overall


reliability of this type make it essential in many instances.

1 t
"0" Ring seal

~-4r-_"--"O_"-"R I n 9 sea I

+-_--t*'~ __~!+--'-Po"'l_'_ethylene perforoted pipe

o
0,
N

_---".j'---'C~o"'a r S e 5 Q nd

r*,~~~+-,-P~e,-,rforated galvanized steel tube

__ ~tvonlzed adapter

.~~it-- __B__o1tom layer (.and)

+-~~~~ ____O.15 m ~_ _ _ -+

Figure 14. Schematic in~ake of a foundation hydraulic piezo-


meter

In Portugal, the electric vibrating wire piezometers are


also used. This type consists of a tip with a diaphragm that
is deflected by the pore pressure against one face. Deflection
of the diaphragm is proportional to pressure, and is measured
by means of various electric transducers. Such devices are
extremely sensitive, and are therefore suitable for
installation in quite impervious and highly plastic clayey
materials.
When fast response or long lead lengths are needed, electric
sensors are used. Though better materials and techniques are
now available for protection, the electric system is still
vulnerable to electric storms, leakage through seals and
breaking of cables. This means that the useful life of this
kind of device is very short (at most a few years). Near these
cells standpipe piezometers are therefore frequently installed
in order to ensure continuity of observations.
Pneumatic piezometers use the diaphragm to operate a control
valve that indicates when applied air pressure is equal to the
water pressure in the filter. As with the hydraulic
piezometer, two tubes are used to connect it to the instrument
house or manhole. Pneumatic piezometers have been extensively
used for dams in North and South America.

4.8. SEEPAGE MONITORING

Knowledge of water seeping through a dam is essential in


assessing its behaviour. First indications of dangerous
deterioration are often given by a change observed in the
seepage rate, as is mentioned several times in the ICOLD
Report on Deteriorations in Dams and Reservoirs (ICOLD, 1979).
Equipment for measuring seepage flows formed some of the
earliest instrumentation installed at dam sites. Where under-
drains or relief wells are connected to manholes, flows are
often measured with v-notch weirs or, when rates are low, by
temporarily diverting the flow into containers of known
volume. More sophisticated equipment includes remote reading
level sensors for recording water levels and hence flows over
weirs, and flow meters that can be either read directly or
arranged with electrical connections so that they can be read
remotely or recorded. The scheme of a v-notch weir placed in
the drainage gallery of Meimoa Dam is shown in Fig. 15.
Normally the measured flows are affected by rainfall. As it
is possible, the readings should be taken when the reservoir
is at top water level and at low water levels in a dry period.
Sometimes, the monitoring scheme is conveniently planned to
measure the flows from the dam and from the rainfall in
different measuring devices.
Besides the flow, it is important to observe the aspect of
the water. Turbid water may be carrying soil particles and can
signify internal erosion. Chemical analysis can reveal whether
the sample contains suspended particles of sailor whether
there are dissolved minerals (Veiga Pinto et. al., 1991).
Recently research has been developed in Sweden (Cederstrom,
1990) for measuring of leakage in the dam by measurement of
the watertemperature in drilled holes and the changes of the
temperature over the year. This method may also indicate if
faults occur by changes in the leakage.
~ w
V'---'N"-'O"-'T'--'C'-'-H'--ll----//
WEIR

CROSS SECTION OF GALLERY PLAN

" D:".
PERSPEX TUBE
6
, D
~
,," 0

I :1
B_R!<SS PLATE
SCALE

..
11

" "'D FLOOR

~: 0 ."

'"
".
.
C>

SECTION A-A' SECTION 8-8'

Figure 15. V-notch weir placed in the drainage gallery of


Meimoa Dam
497

4.9. EARTHQUAKE EFFECT MONITORING

Design calculations for dams in seismic areas include


allowance for the maximum expected earthquake, and sensitive
instruments must therefore be installed for measuring dynamic
effects.
Because the inertia forces developed by earthquake shock
increase with the height of a dam, it may be argued that
modern higher dams will be more liable to damage.
Seismographs should be located in the vicinity of the base
and/or abutment and the crest of the dam, for recording the
ground motion and the response motion, respectively.
Instrumentation of major dam in a seismic zone cannot be
complete without strong motion seismographs for recording
motion and acceleration in three dimensions at positions on
the crest and other parts of the dam. They will record induced
seismicity, explain irregular values recorded by the other
instruments and build up a body of experience on dam behaviour
under earthquake conditions.
Detailed information about seismic effect monitoring can be
obtained in two reports entitled "Seismic instrumentation in
dams" (USCOLD, 1975) and "Guidelines for selection and
installation of strong-motion instrumentation" (USCOLD, 1986).

5. Readings, Processing and Analysis of Results

Effective analysis of the safety of a dam can only be made


from detailed knowledge of dam behaviour, much of which is
provided by readings and processing of results from suitable
instruments.
This analysis can be divided into two levels as is shown in
Fig. 16 (Silva Gomes, 1982). In the first, fast data analysis
is obtained, including data acquisition, data transmission
between the dam and a processing center, data processing and
storage of the corresponding physical quantities, and data
analysis. In the second level, the behaviour of the dam is
pictured by the technicians responsible for safety control,
this being obtained from analysis of results and visual
inspection.
Given below are comments on the activities of safety
analysis based on Portuguese experience in observation of fill
dams.
5.1. DATA COLLECTION

It is widely accepted that in safety control it is very


important to get rapid data processing and a picture of the
dam's behaviour. Nowadays, this is possible by installing a
computer system for processing and storage the monitoring
data. Very recently, a system of this type was developed at
LNEC.
49H

i!:=====:!J
CHECKING
If---~IMEASUREMENTS If---~I RECORDS 1------,
DATA
PROCESSINGI~--~ITRANSMISSIONI~-~
COMPUTER

SECOND CORRECTION
MEASUREMENT WITHOUT NEW
MEASUREMENT

GRAPHICAL AND
ICURRENT MANAGEMENT 1 NUMERICAL OUTPUTS

IGRAPHICAL ELEMENTS 1

INUMERICAL ELEMENTS I

VISUAL INSPECTION

REPORT

Figure 16. Performance evaluation of a rockfill dam (Silva


Gomes, 1982)
499

As it was mentioned before, the frequency of measurements is


settled in the monitoring scheme report.
When the equipment is placed during the construction phase,
the measurements start as soon as the construction operation
allow it. Frequency of readings will depend on progress in
construction.
In Portuguese fill dams it is not usual to have a technical
team permanently at the dam for taking readings of the
equipment. A LNEC team is sent from time to time and the
measurements are made manually.
A first checking of the data is performed immediately after
its collection mainly taking into account the previous
evolution.
The automatic data acquisition of some equipment is now
being developed once the automatic system of data processing
and analysis has been developed.

5.2. DATA TRANSMISSION


It is planned in future to use automatic data acquisition of
measurements and to use local minicomputers for processing
some information and afterwards to send it to a host computer
in the processing center of LNEC.

5.3. DATA PROCESSING AND INFORMATION STORAGE


As has been mentioned, a research programme has been launched
for automating these activities.
In the framework of this programme, the reception of data
takes the following course:
a) input of the data in the computer (the record forms
are prepared to facilitate this task);
b) checking the data (this checking aims the detection
of major errors);
c) calculation of the usual physical quantities by
using adequate algorithms;
d) checking of those quantities;
e) storage of information.
All imputed data are compared with two limits Which, in
principle, will not be exceeded in any situation within the
normal behaviour of the dam (taking into account, for
instance, the extreme normal local combination or considering
the admissible range of variation of the corresponding
instruments). This means that if a value occurs outside the
500

interval defined by those two limits, a major error or


disturbance (of the instrument or even of the dam's behaviour)
has occurred. In case that an interruption of computer
processing occurs, rapid intervention is required (for
instance request for repeating readings in the dam site) so
that the situation may be investigated.
Data is stored in files being updated after each new
acquisi tion, in such a way that any group can be easily
extracted, grouped by type of physical quantities
(displacements, pore pressures, stresses, etc.), by dates or
any other category.
These groups can be used by subroutines (for the automatic
drawing of diagrams, calculating results of quantitative
analysis, etc.). Therefore the system has many subroutines
requiring adequate organization of the stored information, and
supplies the necessary elements for interpreting the dam
behaviour.

5.4. DATA PRESENTATION

The results are usually presented in terms of diagrams and


listings. Diagrams allow a first analysis of the evolution of
the physical quanti ties. The listings sometimes regarding
quantitative analysis permits a more comprehensive interpre-
tation of the results.
5.5. PERFORMANCE EVALUATION
As it was mentioned, the first level of analysis is processed
by computer routines. The second level of analysis is develo-
ped by those responsible for safety control and is based on
the information obtained by the management of the results
(diagrams with the evolution of results, diagrams and drawings
representing particular situations: displacements, stresses,
pore pressures, etc.), and also on information obtained by
visual inspections or documents reporting facts concerning the
life of the dam.
In a first analysis, the trends in the evolution of the
physical quantities are obtained and their values compared
with identical situations. Also important is the crossing of
results of different quantities.
In a second level, an overall behaviour of the dam is
compared with the predicted model. This model is developed by
taking into account the material properties, geometrical
characteristics, external connections, previous behaviour of
the dam, etc.
When acceptable agreement between the predicted and observed
values is achieved, it can be considered that the behaviour of
the dam is normal as regards those particular quantities.
Otherwise, a deeper analysis becomes necessary in order to
verify if this fact is due to the model (insufficiency or
501

inadequacy) or to real defective behaviour of the dam.


When a model can predict values of quantities considered as
representative of the overall behaviour of the dam, it can be
said that the dam behaves normally if good agreement is
obtained between the predicted and observed values. Even in
this case, however, situations may occur which only visual
inspection can detect. This may happen because the abnormal
behaviour has not caused disturbances in the physical
quantities analysed. The analysis of results is more or less
accurate depending on the effective representativeness of the
set of physical quanti ties analysed and therefore, should
always be regarded only as part of the diagnostic. In this
context, visual inspection may supply essential information
for the safety assessment envisaged.
Safety analysis of dams is a continuous activity. Frequent
reports are therefore prepared, helped by automatic routines
that are more directly related to the safety of the dam. A
deeper analysis must, however, be made periodically, involving
all the information available. The periodicity of this
analysis depends on the life of the dam. Usually f after
construction, after the first filling of the reservoir, after
the first year of normal operation and also every five or at
least ten years a very detailed report should be prepared.

6. Visual Inspection

The site inspection of a dam is a visual qualitative activity


which permits detection of any superficial changes or signs of
deterioration. The external evidence can provide important
information regarding performance.
It should allow for detection, analysis and recording of the
smallest changes in the structure or its environment by
following a previously established check list to avoid
omissions. This list should contain, among others, the
following main items; general aspect and homogeneity of the
upstream and downstream slopes and riprap, presence of
vegetation on slopes or abutments that may obscure signs of
adverse conditions, visible signs of slipping, sloughing or
slumping of slopes, presence of wet spots on the slopes, faces
or downstream adjacent zones of the dam, signs of deformation
(settlement, fissures, change in alignment of bench marks,
etc.), foundation conditions {water resurgence, piping, sand
boils, cracks, etc.} and excessive water leakage or seepage.
Any of these conditions may indicate a process that, if not
corrected, can lead to failure of the embankment.
The use of instrumentation to monitor the behaviour and to
detect any deteriorating tendencies is now a standard
requirement for all large dams. Nevertheless, visual
inspection has been the most frequent method used in detection
of deterioration.
Detection of some deterioration has not been borne out in
502

instrumentation data, probably owing to insufficient covering


of the fill, incorrect location of instruments, deficiencies
in operation or in the methodologies adopted in processing,
analysis and interpretation of the results.
The visual inspection must be seen as a valuable and
essential complement of a monitoring system irrespective of
how good this system may be.
Later, in Chapter 7, a detailed presentation of statistical
data on dam deterioration will be described (ICOLD, 1979).
However, notable that from the 1105 cases detected, direct
observation was the most efficient method (951 times), which
emphasizes the importance of this task.
Sometimes, some evidence such as fissures, leakage,
anomalous movements in the shells and banks of the reservoir
that can be observed by visual inspection can only be measured
by the instrumentation in a short perfod that precedes an
imminent collapse.
Occasionally opening of cracks may be measured. Plate
tell-tales or pin tell-tales are in common use and are easy to
read. For instance, any movement between the pins is measured
with a calliper. This simple system is less susceptible to
damage or vandalism, and can record changes in crack width to
an accuracy of 0,1 mm (BRE, 1990).
The visual inspection may be processed at two levels. At a
first level, in frequent observations by technicians living
near the dam. At a second level, a more specialized work may
be done by the responsible geotechnical engineer periodically
with the help of photographs, marks, reports, etc.,
interpreted resul~s of the instrumen~ation readings and
reports from inspections of the first type.
The visual inspection may be execu~ed following specified
rules which may be modified and complemented according to the
observed behaviour of the dam.

7. Safety Evaluation Based on Deterioration

7.1. INTRODUCTION
The safety of a dam, i.e., the level of risk, is considered as
the probability of occurrence of deterioration and it depends
on care taken and knowledge duly applied in design and cons-
truction. For instance, in the design stage the engineer deli-
neates the concept of the structure. This corresponds to
defining the value of the defect, that is the possibility of
a deterioration with a low probability of occurrence. A
deterioration corresponds to abnormal behaviour of the dam,
the reservoir banks or the downstream zone. Usually, the
engineer prepares the project by using deterministic methods,
i. e., he considers that the probability of failure of the dam
is to be zero since failure of such a structure has results
that are often catastrophic. Nevertheless, numerous failures
have occurred throughout the history of dam construction,
which has prompted a reflection on the mistakes behind them,
and has thus contributed to improving the techniques and
methods of dam design and construction.
The analysis of deterioration resulting from abnormal
behaviour of dams shows that such behaviour is usually
produced by a complex of causes and mechanisms calling for
great care in interpretation. However, it seems that one of
the most realistic ways to learn from our mistakes is to carry
out statistical analyses with structures of this kind that
have suffered deterioration. In this way, it is possible to
get an idea about the mechanisms leading to abnormal
behaviour.
Next sections present an approach to the analysis of the
safety of rockfill dams through a statistical treatment of
cases of deterioration reported in the literature. As will be
proved, the deformations are the most important quantity
related to deterioration in dams. Having this in mind, some
tables and graphs with compilation of measured strains in
rockfill dams will be presented in order to provide an
auxiliary tool for estimating predictable strains in dams and
helping the calibration of values obtained from other means
(models, etc.)_ For the sake of clarity these elements are
presented for the construction phase, the operation period and
an earthquake event.

7.2. STATISTICAL ANALYSIS

Experience is a very important factor in the design of rock-


fill dams. One of the ways to profit from experience, limited
though it may be, is to analyse the cases of defective stru-
ctural behaviour. In this context, it seems convenient to
present some statistical data taken from a work carried out in
1979 by the Committee on Deterioration of Dams and Reservoirs
of the International Commission on Large Dams. This Committee
launched a worldwide enquiry on deterioration detected in dams
constructed up to 1975, 1105 cases of deterioration having
been found in dams of all types (above 15 m high) in 33
countries. These data correspond to a sampling of 14700 dams
built in those countries which means a percentage of
deterioration of nearly 10%.
The number of failure cases, i.e., total loss of servi-
ceability of the structure, is about 1%.
The number of cases of deterioration in rockfill dams is 90.
In Table 4 the zones where these incidents have occurred are
presented together with their percentage of occurrence.
What is striking is the high incidence of deterioration in
appurtenant works, which, to some extent, means a high margin
of safety in the dam itself and in its foundation. In fact,
should ele whole number of earth and rockfi11 dams be
504

considered, the incidence of deterioration in appurtenant


works will reach about 33%. Also of note is the large number
of incidents in foundations and reservoir banks which points
to the need for careful exploration and survey of the
geological and geotechnical conditions of the work site.

TABLE 4

AFFECTED ZONE (%)


Appurtenant works . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
Main body . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
Foundation ................................... 19
Reservoir banks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

As for the type of incident, the following cases were


obtained, by percentage (Table 5).

TABLE 5

TYPE OF DETERIORATION (%)


Inadequacy of appurtenant works .............. 44
Internal erosion and seepage (foundation) .... 17
Seepage through upstream facing .............. 13
Internal erosion and seepage (clay core) . .... 8
Inadequacy of slope protection ....... ........ 7
Reservoir slope sliding and seepage ......... 6
Slope sliding (main body) .. ................. 3
Differential settlement (foundation) ......... 2

The majority of deterioration was thus found to occur in


appurtenant works and in the foundation, the latter case being
mostly due to piping and seepage. If only deterioration in the
main body is considered (Table 6), at most frequently found
is associated with excessive infiltration through either
upstream facings or clay cores. The small number of incidents
produced by sliding of upstream and downstream slopes should
also be mentioned.
As regards the occurrence of deterioration in the several
periods of life of rockfill dams, an almost identical
distribution through the different periods is found (Table 7).
After the first five years' operation, dams of this type may
be considered to enter a stable period with little
deterioration.
505

TABLE 6

DETERIORATION IN THE MAIN BODY (%)


Upstream facing leakage . . . . . . . . . . . . . . . . . . . . . 43
Internal erosion and seepage (core) ....... 25
Inadequacy of slope protection . . . . . . . . . . . . . . 21
Slope sliding ............................ 11

TABLE 7

DETERIORATION TIME (%)


First five years . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Not available . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Construction ............................ 20
First filling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
After five years . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

TABLE 8

DETERIORATION TIME (MAIN BODY) (%)


First filling .................... 37
Not available . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
First five years . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
After five years . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
Construction ...................... 5

If only accidents reported in the main body are considered,


then most of them occurred at the first filling of the dam
(Table 8). The analysis of this Table and of that concerning
deterioration in the main body (Table 6) confirm how important
is the behaviour of the facing as regards water seepage due to
fissuration or bad performance of joints. However, it should
be stressed that most incidents occurred before the advent of
heavy vibrating rollers, as is nowadays normal practice. It is
also possible to see the importance of the core behaviour as
regards seepage and piping. As is pointed out, these accidents
tend to occur during or after the first filling. Rapid filling
should therefore be avoided, especially in clay-core rockfill
dams when the deformability of the core material and of the
shells is very different.
506

Lastly, it may be said that, in percentage terms, the


quantity of deterioration per number of dams built has
markedly decreased in recent years, which reflects the
important progress achieved both in design methods and in the
construction techniques of rockfill dams. This has prompted an
increasingly wide acceptance of rockfill embankments as a
solution for construction of dams. Fig. 17 pictures the
situation in numerous countries as to the number of rockfill
dams built compared with the total number of dams.

PORTUGAL

S PAl N

AVERAGE OF 13 COUNTRIES (Portugal,Spain,

..
I nd ia, USA,Mex i co, J a pan, Switzerland, B raz i I,
50
Canada, Italy, France, Yugoslavia,Sweden)

<f) <f)
:l: :l: p
i.0 ..-
0 0
/// I
r .
...J LL.
...J 0
-
/
LL. ...J
os:: /
u f-
0 0 30 /
a: >-
/ /
/ /
20 , /
/

/
j1
.

/
10 /

1950 1960 1970 1980 1990

Figure 17. Percentage of rockfill dams built after 1950

As Table 9 shows, although the number of rockfill dams


constructed until 1960 was small, the number of accidents has
some significance.
The high degree of safety displayed by rockfill dams is made
clear by the small number (14) of failures reported until
1975. Moreover, all these failures were produced by defective
discharge and diversion works, no failure having occurred as
a result of unsatisfactory structural behaviour of the main
body or even of its foundation.
S07

TABLE 9

PERIOD OF CONSTRUCTION (%)


Before 1940 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1940 - 1960 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1961 - 1970 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
After 1970 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

7.3. REMEDIAL MEASURES

As it was mentioned, rockfill dams are structures with high


factors of safety as to catastrophic accidents. Remedial
measures that have been taken are therefore often intended
mainly to maximize benefits, for instance by avoiding
appreciable leakage, without raising the question of stru-
ctural safety.
The remedial measures that have been implemented in rockfill
dams, excluding those concerning appurtenant works (Table 10)
will now be analysed.
Among these remedial measures, minor repairs of the upstream
facing, in joints and slabs, and the construction of an
impervious clay blanket were used. There were also some cases
of repair of the central core zone by grouting after the
occurrence of cracks and cavities.
In the dam foundations, treatment mainly aimed at
compensating the misfunctioning of the grout curtain, and for
this purpose downstream relief wells and upstream blankets and
inverted filters were incorporated. As a rule deterioration in
reservoir banks did not merit any special treatment, except in
cases where some surface waterproofing seemed appropriate.
On the basis of the statistical analysis of deterioration
and remedial measures implemented in rockfill dams, a list of
recommendations to be taken into account in the design and
construction of new rockfill dams in order further to improve
the safety of such structures is given below:

1 - the potential movement of the curtain, taking into


account the deformation of the rockfill, should be
carefully estimated;
2 - the shells should be heavily compacted in upstream
facing rockfill dams and lightly compacted in
earth-rockfill dams;
3 - ample and well graded filters and drains should be
designed;
sox
4 - detailed geological and geotechnical investigations
of the foundations should be carried out, in
particular on hydrological aspects and on the
characteristics of discontinuities.
These points are related to the weakest zones and aspects,
as taught by experience of the construction and behaviour of
dams of this type.

TABLE 10. Remedial measures for rockfill dams

ZONE N2 OF CASES
Repair of upstream facing
(joints and slabs) 9

Impervious core and filters repair 7

Main Body Slope protection repair 5

Impervious core construction over


upstream facing 4

Filling in cracks and cavities 3

Grout curtain treatment 6

Inverted filter construction 4


Foundation
Construction of uplift wells downstream 4

Construction of impervious blanket 1

No repair 3
Bank
Reservoir Treatment to ensure watertightness 1

Rockfill placement 1

8. Dam Safety Regulations


As said in Chapter 7, about 10% of the dams constructed have
had deteriorntion and there was about 1% of failures.
In some of these failures significant economic and human
loss occurred.
Although the rockfill dams are the highest existing dams,
509

the accidents involving these structures have, however, been


relatively minor.
For the reasons pointed out above, the safety analysis of a
dam should be entrusted to persons with a high degree of
responsibility and motivation.
The consequences of a dam failure are felt on a regional
scale and it is thus reasonable to have several persons and
entities responsible for the maintenance and safety of these
structures although the Government is the ultimate entity
responsible for safety. To this end, many Governments have
been preparing and implementing regulations and codes relating
to the safety of these works.
In Portugal, at the beginning of 1990 the regulations of
IiDam safety" were approved and implemented.
Some notes on this document are given below.
The Dam safety Regulations apply to dams that present great
or significant human or economic hazards.
Safety control is assigned to an official entity named,
Authority, for Regulation purposes, which will have the
collaboration of LNEC in execution of activities within the
field of its responsibility.
safety control begins in the design phase and will be
developed during the full life of the dam, i. e., during
construction, first filling, operation, and abandonment and
demolition.
The document defines the competence of the public entities
and obligations of the owners.
Special importance is given to monitoring and first filling
plans.
The monitoring plan has to be based on the corresponding
proposal and be implemented, at least in its final phase, with
the collaboration of the various entities involved in the
dam's structural life.
The plan will contain directions about:
1 - visual inspections (frequency of visits, types of
inspection ami qualifications of the agents
concerned, main aspects to be observed and
presentation of reports);
2 - installation and operation of the monitoring scheme
(physical quantities for assessing forces,
structural performance predictions, specifications
of instruments and accessories and their
installation and use, frequency of measurements
during construction, first filling and operation
phases and their adjustment in case of rapid
variation of forces or of exceptional occurrences,
methods of data collection and processing,
communication procedures be followed in case of
detection of anomalous behaviour, qualification of
510

the agents responsible for installation and


operation of the monitoring scheme);

3 - methods recommended for behaviour analysis and


safety evaluation of the works (behaviour models to
be used considering the main hazard predictions and
reports to be prepared).

The first filling is the most critical phase from the point
of view of dam behaviour and the risk involved. Therefore the
Regulations lay down that a plan must be prepared sufficiently
in advance of the beginning of that phase.
For existing dams, adaption to the Regulations must be made
within a period of five years.

9. Strains Observed in RoekfiII Dams

9.1. INTRODUCTION
Monitoring of fill dams is mainly concerned with the analysis
of safety and deformation behaviour. This has the underlying
assumption that there may be some mechanisms that may endanger
the operation of the dam and produce deterioration. One of the
most straightforward ways to foresee what mechanisms may lead
to abnormal behaviour in such structures is to carry out
statistical analysis of the cases where deterioration has been
observed either during construction or in the operation
period.
statistical analysis of deterioration in rockfill dams shows
that most of them occurred in dams with upstream facing
(though more in dams built before 1960), mainly as result of
excessive strains in the rockfill beneath the waterproofing
membrane. This proves the importance of the new construction
techniques as well as of the new methodologies for predicting
the behaviour of embankments constructed with those materials.
Likewise, piping in zoned dams is related to the deformability
of the core, since the fissuration of the clayey material and
hydraulic fracturing phenomena are often caused by the
core-shell interaction.
These facts show that in ~his type of fill dam monitoring of
displacements may be of the utmost importance for analysis of
the mechanisms that may endanger the safety of these
structures. It can also be concluded that the ordinary
criteria used in the analysis of fill dams stability (limit
equilibrium methods) are not very meaningful when applied to
the design of rockfill dams, so that mathematical models
should be increasingly used to determine the order of
magnitude of deformations in these structures.
511

The following paragraphs will deal with the empirical


correlations that can be adopted for predicting strains in
rockfill dams, obtained from the results of observation of
real cases.

9.2. CONSTRUCTION PHASE


Beliche (Portugal) and Keddara (Algeria) dams were the object
of analyses at LNEC for prediction their structural behaviour
during construction and first filling. These were type A
predictions I i . e. I analyses prior to the construction of
those structures. By studying the behaviour observed in
similar dams the aim was thus to investigate if the order of
magnitude of the displacements calculated in the two dams for
the construction phase was close to the values obtained in
real cases.
The diagram of Fig. 18 presents the strains observed and
those calculated in rockfill dams with central core.

o E FOR M A B I LIT V MOD U L U S (k 9 f /crrfJ TO A 0 AM OF 100m HI G H

1000 500 350 250

H
1m)

280
THERI CHICOASEN
(INDIA)'" -(MEXICO)

240 MICA. OROVILLE


(CANADA) (USA)

200 NEW MELONES


DARTMOUTH A (USA)
-(AUSTRALIA)

IN FERNILLO TALBINGO
160
(MEXICO) LA ANCOSTURA (AUSTRALIA)
(M EXICD)
MAL PA SO .SVARTEVANN
(MEXICO) (NORWAY)
120
LLYN BRIANNE
KEDDARAA
ARGELlA (UNIT~D KINGDOM~ DALECISE (CHECHOSLOVAKIA)

.
MIRA. TERAUCHI
80 SCAMMONDEN (PORTUGAL) (J APAN)
(UNITED KINGDOM)-
.MAUTHAUS KISENYAMA(JAPAN)
LE SU (WEST GERMANY)
(RUMANIA) MON TE DA ROCHA
40 ALVITO BELICHE (PORTUGAL)
(PORTUGAL) (PORTUGAL)

.. CALCULATED VALUES

DEFORM AllON 1%)

Figure 18. Settlements values corresponding to the construc-


tion phase and at the mid-height of earth rockfill dams (some
measured and some calculated)
512

The unit strain is obtained in the lower half of the dam,i.


e., by dividing the settlement of the central zone of the dam
by half its height.
From this diagram, it may be concluded that the unit strain
varies from 1 to 3%, as experience with a large number of dams
teaches. It will be noted that the smallest strains were
obtained in the highest dams. For a hypothetical dam, 100 m
high, the overall deformability modulus is found to vary
from 30 to 100 MN/m2.

9.3. AFTER CONSTRUCTION

After the construction phase, fill dams undergo settlements


due to creep, consolidation, and collapse of rockfill
materials. Collapse strains are instantaneous strains due to
the presence of water.
In a way similar to the analysis already mentioned for the
construction phase, several authors (Soydemir and Kjaernsli,
1979; Clements, 1984; Justo Alpanes and Gonzalez Martinez,
1986 and Dascal, 1987) have done studies on the strains
observed during the lifetime of rockfill dams.
This information is very useful, mainly for the designers of
dams, because it is possible to predict the order of magnitude
of long-term displacements in dams of this type.
Those in charge of the safety of a given dam should be aware
of possible problems if excessive settlements occur in these
structures as compared with those experienced in similar dams.
Prediction of long-term settlements will also be most helpful
for the definition of the camber to be established in the
design stage.
Charts shown in Figs. 19 and 20 produced by Clements (op.
cit.) are based on the results of the observation of 68
rockfill dams, and correspond to the envelopes of the
settlement curves and deflection curves of the crest, in
percentile terms, by considering unit height. The first
reading after construction corresponds to the initial value of
time, and long-term strains were found to evolve with time
roughly following an exponential law.
The empirical correlations of Figs. 19 and 20 show a great
scattering of values, which to some extent reflects the large
number of factors that influence strains in rockfill dams,
such as: foundation characteristics, properties of materials,
construction techniques, etc. In order to predict settlements
in a real dam, the engineer must therefore weigh up results in
the light of his experience, through a careful analysis of the
influence of those factors.
Fig. 19 shows that clay core compacted rockfill dams have
presented, after five years of operation, crest settlements
from 0.2% to 1.2%. Nevertheless, in upstream facing compacted
rockfill dams the same settlements are much lower, about 6
times less than the values obtained in earth-rockfill dams.
513

t (MONTHS)

--
o 2 5 10 20 50 100 200 500

-;.
~~ r-.-- t--....
LB MFD (CR)
........ I"'"
I

\ "\f'-..
LB_SCD
I- ~~ LB_CCD

"
::I: 0.25
l!)
\ UB_MFD (CR)

1\ "'''
UJ
"- ~

"\ ""-
::I:

z-
I-
~

" '\
"
~
0.50

\ "LB_M~D
~
(DR)
a: ~

'" r\
UJ
c..
0.75 \ ~
\
I-
z
UJ
~ ~
\
UJ
...J '\
l\.
l-
I-
1.00

\ \
UJ
III
~
\U~_MFDI(DR) \ UB-lcD I\UB_~CD
1.25

LB _ LOWER BOUND eeD _ CENTRAL CORE DAM

UB - UPPER BOUND eR - COMPACTED ROCKFILL

MFD _ MEMBRANE FACED DAM DR _DUMPED ROCKFILL

seD _ SLOPING CORE DAM

~igure 19. Settlement per unit height of rockfill dams (Cle-


ments, 1984)

It is also worth mentioning the considerable improvement in


the behaviour of rockfill dams with upstream facing, when
these structures started being compacted with vibrating
rollers, which shows that new building techniques may bring
about important improvements in safety as well as in the
economy of this type of structure.
Justo Alpanez and Gonzalez Martinez (1986) studying concrete
facing rockfill dams, obtained interesting conclusions about
the relationship between displacements and leakage. Those
authors pointed out a empirical relationship of the type:

Q 0.297 {,1.562 (1)


max.
514

whe:::-e Q (l/s) is the seepage discharge and 0max (cm) the


maXlmum crest settlement. Thus it can be expected that for li max
< 5 cm very low seepage 10 lis) will occur. For limu > 12 cm
the values of Q may already be important (> 100 lis) and for
those settlements higher than 40 cm seepage may even be disas-
trous (> 1000 lis).

t (MONTHS)

o 2 5 10 20 50 100 200 LB_MFD (CR)

. I\~ T-'
-I-+---\\r+--~~"'---f----+--~"'"""""",,~-~,- - - - + - - - l - - UB_ MFD (C R) -
>--
::r:
1............1 'r-...
-'-'
lJJ
0.25
::r:
>--
Z

1\
::J 0.50
0:
~
:\ , 1"'1
lJJ I
a..
z ~ i

J+.---+----+-----1
0 0.75 ! 1 t----~-+------t-I-~1---1
+-- __
>--
u I \ ' _
lJJ
..J I I !\ I i . U B_S C 0 i

\----+-,---~-- --+-----l---+----i--
1

L
1.00 +---+-+---+---t-
1L

\1
i ---\-
lJJ
0
I I I j __ I
1\
i

1. 25
I

i
I I UB_MFD (DR) i I I

LB - LOWER BOUND CCO-CENTRAL CORE DAM

us - UPPER BOUND C R - COMPACTED ROCKFILL

MFO _ MEMBRANE FACED DAM OR _DUMPED ROCKFILL

SCD _ SLOPING CORE DAM

Figure 20. Def:ieccion pe.r: Jnic lleLght of rockfill dams (Cle-


ments, 1984)
515

9.4. AFTER EARTHQUAKES

Another type of displacement that may occur in rockfill dams


is due to earthquakes.
The analysis of the seismic actions in embankments include
many simplifying assumptions, although recently new advanced
numerical techniques have been used.
For evaluating the dynamic performance of rockfill dams the
observed behaviour of existing dams is therefore the most
reliable method.
A large number of rockfill dams have been subjected to
strong earthquakes in seismic zones (Central and South America
and Japan). Some of them have been subj ected to seismic
loadings that induced peak ground accelerations of 0.1 g or
more at the base of the dam.
Displacements observed in some rockfill dams due to the
effect of seismic loadings are presented in Table 11 (Bureau
et al., 1985).
The observed performance suggests that rockfill dams have
high resistance to seismic loads. It may be taken into account
that the dams referred to in Table 11 were not designed to
modern earthquake-resistant practice and up-to-date constru-
ction techniques. For example, in some of them the rockfill of
the shells was dumped and modern heavy vibratory rollers were
not used.
In Cogoti dam, a 84 m high dumped rockfill dam, a displa-
cement of 28.1 cm was observed after the 1943 earthquake.
Even in this case, only minor rock slides along the downstream
slope occurred, but no face cracks and no increased leakage
occurred as a result of the earthquake.
A simplified method for forecasting the displacements of
rockfill dams under seismic loadings was presented by Bureau
et al. (op. cit.). These authors def ined a parameter, ESI
(Earthquake severity Index) as follows:

ESI A(M-4.5)3 (2)

where A is the peak ground acceleration in g's and M is the


magnitude.
In Fig. 21, vertical settlements observed are compared with
ESI of several dams and the authors have pointed out that
relationship as a simplified method for forecastinq the
516

settlements of rockfill dams under seismic loadings.

10.0

.....
z
UJ
~
UJ
~ 1.0
.....
UJ
III
..J

u
.....
Q:
UJ
> 0.1
UJ
> Observed average
.....
relationship
..J
UJ
Q:

0.01


O.OO1+------+-------t-------t--
0.1 1.0 10.0 100.0

EARTHQUAKE SEVERITY INDEX (ESI)

.figure 21. Relationship between vertical settlements and ESI


parameter (Bureau et al., 1985)
517

TABLE 11. Seismic displacements of rockfill dams

SLOPE EARTHQUAKE CREST


DAM H/V MOVEMEN.
COUNTRY
YEAR COMPLETED UPST. iI YEAR ! M !i R PGA VERT.
H ( m) DOWNST. I ,I (km) PCA HORIZ.
(g) (cm)

I>1ALPASSO
PERU 0.5 1938 6 - 0.10 7.6
1936 0.3 - 5.1
78

COGOTI
CHILE 1.6 1943 8 16 0.20 28.1
1939 1.8 0.38 -
84

MIBORO I
JAPAN 2.5 I 1961 7 16 0.20 3.0
1960 1.8 I I ! - I
! 5.0
131

OROVILLE
USA 2.8 1975 6 6.9 0.10 0.9
1968 2.0 0.12
235

EL INFERNILLO
MEXICO 1.8 1979 8 110 0.12 13 .0
1964 1.8 0.35 4.5
148

LA VILLITA
NEXICO 2.5 1979 8 110 4.5
1967 2.5 0.36 3.0
60

LEROY ANDERSON I
USA ! 2.0 1984 6 16 0.41 1.5
1960 ! 2.0 0.63 0.9
72
i
H - Height
H/V - Horizontal/Vertical
M - Magnitude
R - Epicentral distance
PGA - Peak ground acceleration
PCA - Peak crest acceleration
51~

10. Conclusions
This paper was concerned with the issue of monitoring and
safety analysis of rockfill dams. The main conclusions drawn
up are as follows:

1 - The safety analysis of fill dams is one of the most im-


portant tasks in geotechnical engineering owing to risks
that deterioration may involve.

2 - The analysis of rockfill dams behaviour must be based on


readings of the monitoring devices and a visual
observation judiciously appraised by a geotechnical
engineer with suitable experience in design and
construction of these structures.
3 - The modern equipments for laboratory tests and the mathe-
matical models are indispensable tools for the analysis
of phenomena involved in rockfill dam behaviour.

4 - Monitoring plans should be established on basis of risk


factors associated with the structure.

5 - Regardless of hydraulic aspects, defective behaviour of


rockfill dams are mainly related to their sealing device.
In this connexion, safety assessment is chiefly based on
measurements of displacements and seepage. It is
therefore easily understood that monitoring plans may
present significant differences depending on whether
sealing systems consist of upstream facing solutions or
clay core solutions.
6 - Particular attention should be paid to the analysis of
the behaviour of the dam during the first filling phase,
filling stages being controlled and defined "a priori"
whenever possible.
7 - A good monitoring system is often the result of the im-
portance and care given to the preparation of the
moni toring plan. Observations carried out during the
construction of the work are most useful and may
sometimes bring about changes in the monitoring plan.
8 - In general, the simplest monitoring equipment have led to
the best results. Thus it is usually preferable to adopt
519

hydraulic, pneumatic and mechanical devices instead of


the highly sophisticaded ones, usually of electric type.
Nevertheless research for development of equipment mainly
concerned on this area given the tendency to automation
of readings supported by computer means.

9 - The visual inspection has been the most frequent method


used in detection of deterioration.

10 - Deterioration of rockfill dams is chiefly related to de-


fective hydraulic appurtenant works. Incidents detected
in the central body of rockfill dams are normally
associated with excessive deformation and leakage.

11 - Incidents due to sliding of rockfill slopes are practi-


cally nUll.

12 - On earth-rockfill dams, maximum unit strains of 1 to 3%


have been measured.

13 - Clay core compacted rockfill dams have presented, after


five years of operation, crest settlements from 0.2 to
1.2%.

14 - In upstream facing compacted rockfill dams the settle-


ments are much lower than in earth-rockfill dams, about
6 times less.

15 - The observed performance suggests that rockfill dams have


high resistance to seismic loads.
REFERENCES

BRE (1990) "An engineering guide to the safety of


embankment dams in the united Kingdom", Building Research
Establishment Report, 1-155.
BUREAU/ G.; VOLPE/ R.i ROTH/ W. and UKADA/ T. (1985)
"Seismic design of concrete face rockfill dams". In Concrete
Face Rockfill Dams-Design/ Construction and Performance/
Pub. ASCE, 479-508.
CASACA/ J. (1987) - "Local geodetical networks for measuring
displacements in fill dams"(in Portuguese). Geotecnia No.
50, Lisbon, 67-82.

CEDESTROM/ M. (1990) -"Personal communication", in Advances


Rockfill Structures/ Lisbon.
CLEMENTS (1984) - "Post-construction deformation of rockfill
dams". Proc. ASCE, Journal of Geot. Eng. Division Vol. 110,
N7, July, 573-584.
DASCAL (1987) "Postconstruction deformations of rockfill
dams". Proc. ASCE, Journal of Geot. Eng. Division Vol. 113,
N" 1, Jan., 46-59.
DIBIAGIO, E.; MYRVOLL, F.; VALSTAD, T. and HANSTEEN, H.
(1982) "Field instrumentation, observations and
performance evaluation for the Svartevann Dam". Proc. of the
14th Int. Congo on Large Dams, 789-826.

ICOLD (1979) "Deterioration cases collected and their


preliminary assessment". Committee on Deterioration of Dams
and Reservoirs, 2 Vol.
ICOLD (1982) - IIAutomated observation for the safety control
of dams". Bulletin 41, 1-120.
JUSTO ALPANES, J. and GONZALEZ MARTINEZ / A. (1986) - " Las
presas de escollera com pantalla de hormingon armado".
Revista de Obras Publicas, Marzo, 173-194.
521

SILVA GOMES (1982) IIAutomated monitoring tasks in


Portuguese dams. State of the art and prospects". Proc. 14th
ICOLD, 573-584.
SILVA GOMES I A. and VEIGA PINTO, A. (1986) - "Meimoa Dam.
First filling plan" (in Portuguese). LNEC, Internal Report,
1-17.
SILVA GOMES, A.; VEIGA PINTO, A.; TAVARES CARDOSO, E. and
ALI'1EIDA GARRETT, A. (1987) "Measurement of internal
displacements in fill dams"(in Portuguese), Ibero-American
Conf. on Hydraulic Structures Developments, Lisbon, Vol 1,
151-156.
SILVEIRA, A.: FLORENTINO, C.: MARANHA das NEVES, E.; SILVA
GOMES, A. and PITEIRA GOMES, J. (1984) - "Monitoring dams
according to risk factors". Proc. of the Int. Conf. on
Safety of Dams, Vol. I, Coimbra, 221-226.
SOYDEMIR, C. and KJAERNSLI, B. (1979) "Deformation of
membrane-faced rockfill dams". Proc. 7th Eur. Conf. on Soil
Mech. and Found. Eng., Vol. 3, Brighton, 281-284.
EMILIO, F. T. (1987) "Measurement of horizontal displa-
cements of fill dams - a specific device" (in Portuguese),
Ibero-American Conf. on Hydraulic Structures Developments,
Lisbon, Vol. I, 105-110.

USCOLD (1975) - "Seismic instrumentation in dams", American


Society of civil Engineers, New York.

USCOLD (1986) - "Guidelines for selection and installation


of strong-motion instrumentation", American Society of Civil
Engineers, New York.
VEIGA PINTO, A. (1983aj - "Beliche Dam. Monitoring scheme"
(in Portuguese). LNEC, Internal Report, Lisbon, 1-12.

VEIGA PINTO, A. (1983b) - "Structural behaviour prediction


of rockfill dams" (in Portuguese). LNEC thesis, Lisbon, 1-
-157.

VEIGA PINTO, A. and MARANHA DAS NEVES, E. (1985)


"Prediction of Beliche Dam behaviour during reservoir
filling". Proc. of the 11th Int. Conf. on Soil Mech. and
Found. Eng., San Francisco, 2021-2024.
VEIGA PINTO, A. and SILVA GOMES, A. (1986) - "Beliche Dam.
First filling plan" (in Portuguese). LNEC, Internal Report,
1-17.
522

VEIGA PINTO, A. and SILVA GOMES, A. (1988) "Apartadura


Dam. Monitoring scheme" (in Portuguese). LNEC, Internal
Report, 1-23.

VEIGA PINTO, A.; MATIAS RAMOS, C.; MARANHA das NEVFS, E. and
OLIVEIRA LEMOS, F. (1988) "Beliche Dam. structural
behaviour and safety in December 1988" (in Portuguese).
LNEC, Internal Report, 1-42.

VEIGA PINTO, Ai QUINTELA, A.i SILVA GOMES, A. and COELHO, A.


M. (1991) - IoBeliche Dam. Study of a foundation leakage".
Proc. of the 17th Int. Congo on Large Dams, Vienna.
CHAPTER 17
PR INCIPLES OF ROC KFI LL HYDR AU LI CS
R . MARTINS

1. INTRODUCTION

1.1. Definition of rockfill hydraulics

What should be understood by the term "roc k fi l l hydraulics " ?


Obviously it is neces sary to begin by defin in g " rockfi ll ".
There is not , as is kno wn , any single definition . It
varies from author to author , varies according to the point
of view and it has varied with technical evo lution (meaning
t hat t oday much smaller particles are considered as part of
rockfi ll than was formerly the case )
From a hydraulic po int of view , very small particles cannot
be considered as part of roc k fi ll . The behaviour of th ese
particles under flow action , as regards friction head losses
(open channel flow) , stability or permeability , is a subjec t
belonging to classical hydraulics in connection with soil me-
chanics , and does not form part of the spe cific behaviour of
roc k fill . This subject reminds us of the titles of chapters
such as " flow resistance ", " sediment transport " and
" groundwater ", and the names of scientists such as Moody,
Shields and Darcy .
However , rockfill hydraulics should be extended t o sma ll er
particles than those associated with the word " rockfill " wh ic h
are rock blocks with a minimum dimension of , say , 10 cm .
What, the n, is the l imit?
There is no strict limit , dictated by completely objective
criteria. It is rather a nominal limi t that it is possible to
fi x at 1 cm on th e basis o f some practical considerations and
indications of some authors such as Veiga Pinto , 79 .
Rockfill hydrau lics will the r efore be the study of the inter-
action bet ween f lm. and irregular , loose particles , with a
characteristic dimension larger than 1 cm(1) . Natura l ly r oc k-
fill hydraulics has an interfac e with the hydraulics of
porous media with smaller particles . The fact that the three

(I) sometimes 2m is considered the exceptional maximum


dimension and 60cm the usua l maximum dimension (a typical
range being 20cm-60cm)
523
E. MarulI/JII liM Nr"rs (rd,). Ad,wltcs illl/(J('~fill Sf,."..(",..'s. 523- 570.
Cl l 99 1 KII,.,."r Academic Puhlishers.
524

names quoted, Moody, Shields and Darcy, always appear when


studying rockfill hydraulics, is proof of that.

1.2. Complements to the former definition

Irregular shape refers to the blocks as they are found in na-


ture or obtained from the quarry. But some principles of
rockfill hydraulics are pertinent to regular-shaped blocks,
such as those employed in coastal engineering.
Rockfill is, as a rule, a loose material. Thus, blocks con-
nected by some kind of cement are not rockfill. It should be
noted, however, that modern compacted rockfill is not a loose
material - but, precisely for that reason, the principles of
rockfill hydraulics are not, in general, adequate to this ma-
terial. Anyway it is possible to employ some principles of
rockfill hydraulics in two special cases of non-loose materi-
als: gabions and self-spillway rockfill dams with anchor bars
or wire mesh.

1.3. Scope of rockfill hydraulics

A complete monograph on rockfill hydraulics should be divided


into four parts:

characterization of rockfill
general principles
analysis of the different rockfill structures that
involve rockfill hydraulics
special matters.

None of these parts - namely the first - should overlap the


treatment of the theme from the geotechnical point of view,
which is, obviously, the main point of view. Rockfill hy-
draulics is an appendix, although sometimes an important ap-
pendix, to the more general subject rockfill.
The general principles of rockfill hydraulics concern three
points already referred: friction head losses, stability un-
der flow and seepage flow.
A list of works in which flow-rockfill interaction may oc-
cur is:

non-conventional rockfill dams (with throughflow or


overflow)
hazard scenarios related to conventional rockfill dams
rockfill cofferdams
river closures
rockfill protections in several types of works (in flu-
vial hydraulics, in structures related to dams and
transportation networks, etc.)
breakwaters
porous dikes designed as fish barriers.
525

Parkin analyses in Chapter 18 the first of these works.


The fourth part of this monograph should include five spe-
cial points of unequal importance:

protective filters
protection of the upstream slope of earth dams
hydraulics of natural porous media with high permeabil-
ity, for instance highly fractured rock masses
gabions
hydraulic physical modelling of rockfill.

Brauns analyses in Chapter 10 the first point.Filters can


protect rockfill works and can be made of materials analogous
to rockfill.
The second point is a special point owing to the type of
flow in a non-maritime work.
The third point has the special feature of dealing with
natural media. It therefore belongs more to the field of ge-
ology than to the civil engineering. It should be observed,
by the way, that, besides these two fields, others may in-
clude the study of hydraulics of porous media. Among them
Scheidegger, 60, indicates petroleum engineering, water pu-
rification, industrial filtration, ceramic engineering and
powder metallurgy.
The special character of the fourth and fifth points is ob-
vious. (1)

1.4. Subjects dealt with in this chapter

We will treat those that correspond to the first and second


part of the above mentioned monograph, i.e.:

characterization of rockfill
friction head losses in open channels
stability of rockfill under flow
seepage flow.

These matters will be analysed in a practical and simple


way. Needless to say, these adjectives are not flattering but
objective, resulting from the interaction between character-
istics of the subject and characteristics of the author. The-
ory and complexity have an outstanding role in civil engi-
neering.

(1) it seems that hydraulic transport in pipelines should not


be included among these points because dimensions and unit
costs of rockfill make this process improbable
526

2. CHARACTERIZATION OF ROCKFILL

2.1 Preliminary hypotheses

From the point of view of rockfill hydraulics(l) it is usual


to formulate the following main hypotheses:

undeformability of blocks
in a given set of blocks the position of each block in
relation to the others is constant
the material that forms the blocks is impervious
the roughness of this material (surface roughness) is
not relevant when compared with shape roughness.

None of these hypotheses is totally confirmed. In fact, one


of them, the last, will be the object of reserve in this
text. Furthermore, there are supplementary hypotheses, con-
cerning each of the three parts into which the general sub-
ject is divided. Moreover, there are other sources of uncer-
tainty, perhaps more important, that will appear mainly in
Sections 2.2, 2.3 and 2.4. As was said by Parkin, 63: "a cer-
tain randomness of results is inevitable in dealing with
macroscopic particles". Probably it would better to replace
"a certain" by "a great".
If these main hypotheses are accepted, a given set of
blocks would be completely defined if:

dimension and shape of each block


relative positions of the blocks
specific gravity of each block

were known.

This knowledge is obviously unattainable and this is one of


the origins of Parkin's remark. In a correlated field, sedi-
ment transport, analogous reasons have led to analogous re-
marks.
However, an approximation of this knowledge is possible and
this is dealt with in Sections 2.2 to 2.6.

2.2. Size

Let us consider a single block.


Rockfill blocks have a very irregular shape. However, prac-
tice determines that a single dimension shall characterize

(1) it should be noted that, from this point of view,


rockfill is not very compacted, is rather pervious and
stresses among blocks are relatively low
527

the block (owing to what is called characteristic dimension


or, commonly, characteristic diameter) .
The characteristic dimension, d, can be defined in three
ways:

nominal diameter, d n : side of the square passage in a


sieve in which the block passes practically without
clearance (i.e., the block is retained in the next nar-
rower sieve of a given series with side passage dn-l)
mean diameter, d: mean of three dimensions of the
block, dl' d2 and d3' measured according to three or-
thogonal directions (Maranha das Neves and Veiga Pinto,
77 )
equivalent diameter, de: diameter of the sphere whose
volume equals the block volume.

It is, therefore, possible to make d=d n or d=d or d=d e (or,


according to Stephenson, 79, as a first definition variant,
to make d=(d n +d n -l)/2).
The third definition is more precise but less practical,
and so the first and second are more often used.
Let us now consider a set of blocks.
Mean diameter of a set of k blocks, d m, is defined as the
diameter of the mean volume block (or of the mean weight
block, since it is usually considered that all the blocks of
a given set have the same specific gravity).
Considering the same shape for all the blocks of the set,
we have:
(1 )

d 1, d2 , .. ,d k being the characteristic dimensions of the k


blocks.
The most usual way of characterizing the dimension of a set
of particles is the classical sieve analysis. It should be
also observed that this analysis implies the hypothesis of
equal shape for all the particles of the set.
On the basis of the grain-size distribution curve several
parameters, later referred to, can be obtained.Here it can be
noted that dso differs from dm.dSO' as a rule, is larger than
d m, the difference not having, in general, an important mean-
ing.
If the particles of a given set are separated by counting
(and not by sieving) we also get a curve similar to the above
mentioned grain-size distribution curve but it is obvious
that the two curves are different for the same set of parti-
cles.
It is evident that either sieve analysis or counting and
measuring the blocks one by one is, in the case of rockfill,
52!!

a difficult operation, even if we utilize a relatively small


sample.
In some cases, as Leps, 73, suggests, it is possible to use
an expeditious method, not very precise, that consists of es-
timating a dominant diameter for a given set of blocks by a
simple visual inspection, eventually complemented by some
measurements. This dominant diameter will have a value simi-
lar to d m or dso, probably larger. The choice of this inter-
mediate value between average and maximum dimensions depends
on the type of hydraulic problem. Dominant diameter can be,
for instance, adopted in head loss studies.
Two examples of the calculation of these diameters are pre-
sented in Annex I.

2.3. Shape

Blocks can have several shapes. (1)


One possible classification according to Stephenson is:
spherical,rounded, cubical, angular and elongated (Fig. 1).

spherical rounded cubical angular elonualed

Figure 1. Block shape (after Stephenson)

Scott, 63, observes that the most usual shapes of large


particles are the angular and the rounded shapes.
One parameter alone cannot define precisely the shape of an
irregular particle. However, in practice, we consider only
one parameter or we consider sufficient a qualitative de-
scription of the shape. We also consider, as above mentioned,
that all blocks of a given set have the same shape.
The usual definition of shape coefficient, Cs , in the case
of rockfill(2),is:
v
(2)
-3
TId 16

V being the block volume. Other definitions consist of re-


placing d in Eq.2 by the diameter of the sphere that circum-
scribes the block or in making Cs = V/d 3 .

(1) some authors consider that only angular blocks are


rockfill; others distinguish shape from roundness
(2) assuming that sieving is not practical
529

Conceptually, it is not possible to say that anyone of the


definitions is better than the other two, but practice recom-
mends the first (as is indicated by Maranha das Neves and
Veiga Pinto, 77), its meaning being obvious :C s = 1 for a
spherical block. These authors indicate 0.5 to 0.8 as a usual
range of C s values, the first value corresponding to a defi-
nitely angular block and the second to a definitely rounded
block. (1)
Eq. 2 defines the shape coefficient of a single block.
There are two ways of estimating the shape coefficient of a
set of blocks (with a previous choice of a sample): mean of
the shape coefficients of each block or in Eq. 2 making V
equal to the mean volume of the blocks and d equal to d m (or
dsO) .

2.4. Disposition

Void ratio, e, is the main parameter that characterizes rock-


fill disposition (related to porosity, n, by the known equa-
tion n = e/(l+e.
Given the imperviousness hypothesis there are neither pores
nor microfissures, and so the concept of porosity is not am-
biguous.(2)
It is usual to assume the homogeneity and isotropy hypothesis
(in a macroscopic scale). Thus, a single value of e (or of n)
characterizes a given set of blocks. Fig.2, from Bear, 75,
illustrates very clearly this hypothesis.
It should be remarked that the same set of blocks can be
laid in different ways although keeping the same void ratio.
It is therefore understandable to ask if a different be-
haviour under flow corresponds to each disposition. Practice
does not show that (with the exception of salient blocks in
stability problems) and it would be unlikely for it to happen
given the other uncertainties of a higher order of magnitude.
Void ratio depends on block shape, gradation and degree of
compactness.
According to the rockfill hydraulics perspective (angular
or rounded particles, not very far from uniformity and not

(1) the relation between de and d can be obtained on the


basis of these values: de approximates 0.8 d in the first
case and 0.9 d in the second case
(2) besides imperviousness there is no practical interest, in
rockfill hydraulics, in distinguishing between volumetric,
areal and linear porosity, and all pore space can be
considered interconnected
530

very compacted), void ratios do not vary very much and are
rather high.In

these conditions some researchers have obtained void ratios


between 0.5 and 1.2 (void ratio increases with uniformity and
is higher for angular particles than for rounded particles) .

1.0
.....
>-

cr.>
=
=
=
CL..

OL-------------~----~----------
AVERAGING VOLUME SIZE

Figure 2.Porosity of a porous medium as a function of the av-


eraging volume size (after Bear)

The specific surface, sv, will be later in this text relat-


ed to void ratio. Sv is defined as the ratio of the surface
area of an individual particle to its volume. Specific sur-
face depends on block shape. This means that it is a kind of
shape coefficient. But, being dimensional (dimension V 1 ), it
also depends on block dimension, increasing when this de-
creases.
The specific surface of a set of blocks also depends on its
gradation, although this dependence is weak.

2.5. Specific gravity

The specific gravity, s, of the natural materials that form


blocks, has a relatively low variance. This is a supplemen-
tary reason for considering, as we have already mentioned, a
single specific gravity for a given set of blocks.
Fig. 3, according to data provided by Pereira, 85, presents
the specific gravity of several rocks.
531

1.0

Figure 3. Specific gravity of rocks

Usually variation in even lower (2.5 to 2.9 according to


Leps). However, it should be pointed out that relative varia-
tion increases when the blocks are submerged.
The values of specific gravity, given the initial hypothe-
ses, correspond to the values of the apparent dry unit
weight.

2.6. Friction angle

Finally, in this characterization of rockfill, we will refer


to the friction angle, , a parameter that belongs to a bor-
derline with specifically geotechnical areas.
The concept of friction angle,the angle whose tangent is
the ratio of tangential to normal forces in a limit situa-
tion, is applicable to rockfill structures, giving origin to
the more general concept of internal friction angle.
From the point of view of rockfill hydraulics we should
make three observations:

first: the limit situation considered is the situation


in which the motion of the block begins (static
friction angle)
second: as mentioned, normal forces are relatively weak;
in these conditions the angle of repose is a good ap-
proximation of the friction angle (1) (see Fig. 4, from

(1) according to Ulrich, 87 who defines angle of repose as


"the greatest angle to the horizontal that is formed
naturally by the inclined surface of a rubble mound";
obviously it is possible to obtain stability with higher
angles by careful arrangement of the materials; on the other
532

Stephenson), it being usual to consider 400 as a nomi-


nal value of the angle of repose in rockfill
third: cohesion has non-zero values; this is due more
to the interlocking among blocks than to the surface
roughness; thus rockfill is a fractured medium with co-
hesion, presenting an appreciable shear strength (1).

=
=
""" 350 V
I - - - +- \,\\\~~~

./
./
.
~
t\ --
~ \~

~\\\'6\
J.

./
l.oo"
-- --
J..--

~
~
- ...
I"'" \'l.~./
V %\\~;;
~

-
L.,...oo""
~~\
30
0
..... I,..;'"

25"
1 10 100 1000
Stone diameter [mm)

Figure 4. Angle of repose of rockfill (after Stephenson)

2.7. Final comments on sources of uncertainty in


rockfill hydraulics

The choice of the representative grain size of a given set


of rock particles is, perhaps, the main source of uncer-
tainty, and has a certain importance in the three subjects

hand there is no indication of significant modification of


the angle of repose when rockfill is submerged (interlocking
being more important than surface roughness)
(1) it is usual to employ the word strength in relation to
rockfill, although this is not very rigorous, precisely
because rockfill is a fractured medium
533

of this chapter: head losses, stability and seepage.


Block shape is also a general source of uncertainty.The ar-
rangement of blocks has a certain influence on what con-
cerns head losses and stability, it being obvious that, in
these two subjects, dimension and shape do not completely de-
fine the situation: for instance, a regular and compact ar-
rangement diminishes the head losses and increases stabili-
ty.
The influence of the block arrangement is smaller in the
case of seepage flow if we know the void ratio.In fact, as
mentioned in Sect ion 2.4, the hypothesis: a set of blocks
with different arrangements but with the same void ratio has
the same permeability characteristics,has experimental sup-
port.

3. FRICTION HEAD LOSSES IN OPEN CHANNELS

3.1.Preliminary remarks

Determination of friction head losses is a classical problem


in hydraulics.
The usual expressions are based on results obtained with
non-high relative roughnesses (say, up to 0.05)(1). relative
roughness being defined by:

k' (3)
f=
D

where k'is the absolute roughness and D the hydraulic diame-


ter (equal to 4 R, R being the hydraulic radius~)) .
Sometimes relative roughness appears as being the parameter
d/h, in which h is the flow depth. This results from identi-
fying absolute roughness with d (this will be commented on
later) and from taking h instead of hydraulic radius (this is
a usual approximation in hydraulics in the case of wide chan-
nels) .
Obviously fand d/h are related by:

1 d
f (4 )
4 h

(1) according to Moody's diagram that synthetizes the main


data relative to friction head losses; besides this diagram
other authors, for instance Bathurst,78, mention the
occurrence of a blocking effect for high relative roughnesses
that increases resistance
(2) as is known, hydraulic radius is defined as the ratio of
the cross-sectional area of flow to the wetted perimeter
534

In the case of rockfill(1) relative roughness may be rather


higher than 0.05.
Even if E< 0.05 absolute roughness is relatively high and
the flow is no longer uniform.
As is known, uniform flow is a steady flow (i.e., with ve-
locity not dependent on the instant considered) in which ve-
locity is constant along each trajectory. The cross-section
must remain the same in a uniform flow and this does not hap-
pen if the boundary is a rockfill boundary (see Fig. 5).

I
f
f
I
I
Art88~CC'

Figure 5. Flow with rockfill boundary

This question is theoretically relevant since friction head


losses expressions have been based on uniform flows. In prac-

(1) when we are speaking about rockfill we are usually


thinking of artificial channels; some results presented in
Section 3 are, how-ever,applicable to natural river channels
having beds with cobbles and boulders
SJ5

tice, considering other uncertainties, we can use these ex-


pressions in the case of rockfill boundaries.
In fact, in this chapter we will always consider uniform or
quasi-uniform flows. Naturally, in the more general field of
rockfill hydraulics, non-uniform and unsteady flows must be
considered.
Another question that can be asked is: what is the effec-
tive cross-section of flow.
We can measure flow depth in relation to an impervious lay-
er under flow, or in relation to the plane that contains the
block tops, or in relation to any intermediate plane. It
seems less ambiguous to measure depth as indicated in Fig. 5.
In fact, even under conditions favourable to the development
of seepage flow, seepage discharge does not exceed, say, 10%
of the open channel discharge. This uncertainty in measuring
water depth is an additional cause of the dispersion usually
shown by results in rockfill hydraulics.

3.2. Resistance laws

Friction head loss calculations are based on the general ex-


pression (see, for instance, Quintela, 81), known as Darcy-
Weisbach equation:
2
i f 1 U (5 )
D 2g

in which i can be considered the hydraulic gradient (see An-


nex II), f is the friction factor, U is the mean velocity
(equal to Q/S,Q being the discharge and S the cross-sectional
area of flow) and g is the acceleration due to gravity.
Relevant flows in this sect ion are, moreover turbulent,
rough turbulent flows (due to the magnitude of velocities,
cross-section dimensions and roughnesses). Thus f depends
only on relative roughness and is given by the Karman-Prandtl
equation (see, for instance, Quintela):
1 3
if = 2 log E (6)

The constant that originally appears in this expression is


3.7 instead of 3. However, the value 3 is considered more ad-
equate to open channel flow (see, for instance, Henderson,
66) .
Even in the classical case (non-high relative roughnesses),
the joint use of Eqs. 5 and 6, in principle the most precise
method, faces the difficulty of choosing the absolute rough-
ness (and thus E). This difficulty is evident in the case of
rockfill.
536

Purely empirical resistance laws are therefore used. Among


them perhaps the most popular one is the Gauckler-Manning ex-
pression(1):

U = Ks R2/3 i 1/2 (7 )

in whick Ks is a dimensional coefficient (L l / 3 T -1) depen-


dent on the type of solid boundary ~).
It can be noted that resistance depends also on the cross-
section shape (hydraulic radius does not synthetize all shape
properties). In the case of rockfill, given other uncertain-
ties, this dependence does not seem relevant.

3.3. Function f (E) in case of high re~ative roughness

Several authors have sought to adapt Eq.6 to the case of high


relative roughness (among others Gordienko, 67, Thompson and
Campbell, 79 and Graf,89).
This task faces two major difficulties:

the choice of the characteristic dimension


the choice of the absolute roughness/characteristic di-
mension ratio

besides other uncertainty causes, as mentioned in Section


3.1: way of defining cross-section and approximation of h to
R.
As regards the choice of the characteristic dimension una-
nimity exists only on one point: it should be at least a mean
dimension, but between a mean dimension (which can be de-
fined, in fact, differently from Eq.1) and the maximum dimen-
sion, several representative dimensions have been indicated.
Different values are also indicated for the relation be-
tween k' and d, varying from k'= nd to 4.5 d.
It seems that a practical way of overcoming these difficul-
ties might consist in taking as representative dimension d sc
or d m or the dominant diameter (according to Leps), in con-
sidering k'= d (3) and then in plotting on a graph f,E the
available results, from laboratory and from nature, obtained

(1) also known as Strickler expression, Manning expression or


Manning-Strickler expression
(2) strictly speaking it depends on R too; it is this fact
that makes this exprssion, in princple, less adequate
(3) note that this involves the hypothesis already made of
considering surface roughness to be irrelevant
537

with irregular particles (4). According to the indication of


Thompson and Campbell (sustained by Henderson) when the
known block dimension was the median diameter (see Annex I)
the characteristic dimension was taken as being almost twice
that diameter.
The results of Thompson and Campbell and Graf (also of Gor-
dienko) are synthetized in Fig.6.

1.0
f VI
/

I ~V
V
8
ri.==lIDIi. + II~
~
~" 4,
S

~~
,
*0111 - 0.8, J luoll
""~'~<,#'I!,
I

~
~\\

1 /
~
0.01 9 -----., . i
'--high relatIVe roughness
0.05 0.1 0.1
- comsponGing to vt- = Z log +
O-GOROIIKO, 61
*-Similar In RYABOY, 14 results
I "",dIBD I. lKOMPSOI and CAMPBElli

Figure 6. Friction factor in case of high relative roughness

3.4.Data for the use of the Gauckler-Manning expres-


sion

Literature often indicates values of Ks for rockfill (more


precisely for coarse gravel) . The value most usually mentioned
(see, for instance, Chow, 59) is Ks = 30 m1 / 3 /s.
It seems, however, that this value is more adequate for
finer particles than for typical rockfill and, furthermore,

(4) there are other studies on high relative roughness flow,


but since the roughness is obtained with regular shaped
elements, it does not seem adequate to apply these results to
rockfill
538

more adequate for "normal" (i. e., non-high) relative rough-


ness.
The awareness of it not being very logical for Ks not to
depend on particles dimension leads other authors to propose
expressions such as Ks = constant/d 1/6 (1).This tYVe of ex-
pression has been utilized in sediment transport studies (and
thus applied to small particles, e.g. sands) and it was ini-
tially presented by Strickler in 1923 (see Henderson) .
The several authors who have applied this type of expres-
sion to rockfill (for instance Maynord et al., 89) propose
different expressons but all of them are very similar to (m.
sec units) :

20
1/6 (8 )
d

There are different indications as regards the pertinent


characteristic dimension (from dSO to d max ) It is difficult
to give a more precise indication beyond saying that it is
necessary to take dSO or a larger dimension (to take the max-
imum dimension does not seem logical). This is similar to the
use of the dominant diameter concept.
It should be noted that Ks = 30 m 1/3 /s implies, according
to Eq. 8, that d is nearly 9 cm. This agrees with the first
paragraph of this section.
It should be repeated that,as happens with Eqs.S and 6,
also Eqs. 7 and 8 are not adequate for the calculation of
flows with high relative roughness.

3.5. Conclusions

The calculation process of friction head losses in open chan-


nel flows with rockfill boundary begins with the determina-
tion of (2).We can admit O.OS as being the nominal limit that
distinguishes high and "normal" relative roughness (it is
situated between dlh = 0.3, according to Bathurst, and hid =
10, according to other authors) .
If ~ O.OS we can use Eqs.S and 6. An alternative way con-
sists of using Eqs.7 and 8.These last expressions lead to
higher head losses (an increase from 6S to 45% when varies
from 0.01 to O.OS).
If > 0.05 we can use Fig. 6.

(1) the dependence of Ks on the power 1/6 of d is due to


dimensional considerations
(2) in many cases we do not know h in advance, and it is
necessary to use an iterative process
539

4. STABILITY OF ROCKFILL SUBJECT TO FLOW

4.1. Preliminary remarks

Let us consider in this section, as already said,the case of


uniform or quasi-uniform turbulent flows (quasi-uniform flows
are mentioned because, strictly speaking, there is no uniform
flow in horizontal channels or bends). In fact, in general
and from a practical point of view, steady gradually varied
flows or gradually varied unsteady flows can be assimilated
to uniform flows. If the degree of non-uniformity or the de-
gree of unsteadiness is high (for instance, respectively, in
stilling basins and in breakwaters) special questions arise
that are more conveniently treated when specific works are
studied.
The theoretical analysis of the stability of rockfill under
flow is based on the equilibrium that exists in the critical
conditions (incipient motion conditions) between the forces
due to flow and the forces derived from the particle
weight. (1)
This analysis employs two basic expressions.
The first, more related to the case of large particles, is
the classical drag force equation:
2
U
FD = CD'Y A - ( 9)
2g

in which Co is a dimensionless coefficient depending on flow


velocity and particle dimension, 'Y is the unit weight of wa-
ter and A is the particle area projected on a plane normal to
the flow.In the case of rockfill, the development of theoret-
ical analysis faces great difficulties.On one hand turbulence
and the irregular shape and arrangement of the blocks make it
impossible to know CD, the direction of FO (and thus A) and
the appropriate value for U. On the other hand, knowledge of
the resistance forces is complicated by the existence of in-
terlocking among blocks (many situations being possible be-
tween two limit cases: very salient block and block quite in-
tegrated in a layer) .
The second, more related to the case of sediment transport
(giving origin to the classical Shields criterion) is:

'to = 'Y h i (10)

where 'to is the shear stress at the solid boundary.

(1) in the case of rockfill interlocking should also be


considered
540

Note that in this expression h appears,differently from Eq.


9. In fact, in stability expressions, h may figure
(explicitly or implicitly) or not figure.Zarkhov,85, observes
that, in general, flow depth is not a very important parame-
ter for the stability of large particles.
The difficulties found by theoretical analysis give a major
role in this matter to experimental research. Theoretical
analysis, associated with dimensional analysis, has however
produced results, such as the form of expressions and the es-
timation of the effects on stability of non-horizontal beds
and side slopes.
Finally, four remarks:

as is typical of rockfill studies there are differences


of opinion about the representative dimension; usually
a mean dimension, such as d50, is considered (an alter-
native is to consider to be on the safe side a lower
dimension, such as d30)
some authors (for instance Graf) call attention to the
fact that high relative roughness, similarly to the
case of friction head losses, influences stability con-
ditions
seepage flow may produce forces against stability; in
the classical case of linings, relatively thin layers
on a impervious surface, these forces are not important
the definition of the critical conditions, decisive for
obtaining the stability expressions, is not a fully ob-
jective matter.

4.2.Stability in bidimensional channels with horizon-


tal or quasi-horizontal bed and non-high relative
roughness

We begin by presenting results obtained on the basis of Eq.


9. According to the classical work of Izbash on river clo-
sures (see, for instance, Izbash and Khaldre, 59) we have:

(11)

in which Ub is the mean velocity in the vicinity of the block


("bottom velocity" or "velocity against stone") in the criti-
cal conditions and Cr is a dimensionless coefficient. Differ-
ent values for Cr have been presented, according to particu-
lar conditions. One of them, 0.86, is probably an acceptable
value in all conditions.
Neill,73, proposes a "mean" curve d = f (Ub) of four curves
utilized by USA agencies (California Division of Highways,
Bureau of Reclamation, Corps of Engineers and Bureau of Pub-
lic Roads of Washington) .This curve is approximated by (m.sec
54l

units, considering s 2.65, in the range 2 m/ s < Ub < 6


m/s) :
2
d 0.038 Ub - 0.05 (12 )

Precise knowledge of Ub is very difficult. Determination of


velocity profiles, even in the simplest case (uniform flow in
a smooth wide channel), is a open question in hydraulics.
Considering other uncertainties, we think that it is accept-
able to use an expression such as:

0.71
0.68 log hid + 0.71 ( 13)

which is mentioned in the publication of the USA Corps of En-


gineers "Hydraulic Design Criteria".
Zharkov mentions the Rakhmanov expression (applicable to
large particles in the range 0.006 < <0.05):

U = 2.5 Ycid (14)

This expression does not include flow depth, explicitly or


implicitly.
Shields'criterion, presented in 1936, is based on equilib-
rium in the critical conditions between the shear stress (Eq.
10) and the resistance due to the particle weight.In these
conditions, and for the turbulent flows that are typical of
civil engineering, we have(1):

0.056 (15 )
(s - 1) Y d

From this equation we easily get:

U = 0.47 ~2q d (s-l)/f (16 )

f can be calculated by Eq.6.


Finally Maynord et al.present an expression that is a vari-
ant of one presented by the first author in 1978:

2.5
~ = 0.16 ( ~ ) (17)

(1) the experimental coefficient 0.056 (in Eq. 15) is


sometimes presented with slightly different values
542

4.3. Case of high relative roughness

Let us accept = 0.05 as the nominal limit, already men-


tioned, that separates high relative roughnesses from the
"normal"ones.
What are the consequences for the stability of rockfill un-
der flow of a situation of high relative roughness?
It can be said that there are three consequences:

U become a better approximation of Ub


'to increases (because i increases - see Section 3.3),
this being unfavourable to stability
Shields' parameter, i. e., the ratio between actions
and resistances in the critical conditions, increases
too, this being favourable to stability.

Among the five expressions mentioned (Eqs. 11, 12, 14, 16


and 17) we consider that the Izbash expression is the best,
since it was obtained in a context of high relative rough-
nesses (river closure) .
Shields'criterion would be applicable if the data relative
to the value of the Shields' parameter in this situation
showed agreement. But this does not happen: for instance Wang
and Shen,85, indicate the value 0.25 and Graf data indicate a
value of, say, 0.09.
However there is other pertinent expression. It is the
Olivier expression, presented in 1967, based on laboratory
tests in the following situation: overflow dam, with down-
stream slope relatively gentle and protected with rockfill.
It is a typical situation for the occurrence of high relative
roughness flow.
The Olivier expression, as is mentioned by Thomas, 76 is
(m. sec units) :

(18 )

in which q is the discharge per unit width (critical condi-


tions) and a the angle of the downstream slope to the hori-
zontal.
If we admit the establishment of uniform flow on the down-
stream slope we easily get (s = 2.65, and m. sec units):
0.45 C.8S
3.82 d h (19 )
U
O.3S
f
543

4.4. Channels with non-horizontal bed

In a non-horizontal channel (positive slope) the critical ve-


locity should be modified according to:

Uc:x. = ec:x. U (20)

in which Uc:x. is the mean critical velocity in a channel with


bed slope tan c:x. and ec:x. is a coefficient.
Simple theoretical considerations lead to:

ec:x. = V cos c:x. -


sin c:x.
tan <!>
(21)

Ulrich, considering an alternative theoretical model, pro-


poses:

ec:x. I~os c:x. _ s sin c:x.


'V (22)
=A

s-l tan <!>

and, besides this, that the angle of repose be replaced by


the bearing angle with a nominal value of 75 0 (the idea is to
replace the angle at which begins the motion of a block Qfi a
layer, by the minimum angle at which a block in a layer fall
out) .
The following table quantifies these corrections (s=2.65):

tan (
0.01 0.03 0.05 0.07 0.09 0.11 o .l3 0.15
<!>

40 0 0.99 0.98 0.97 0.96 0.94 0.93 0.92 0.90


Eq. 21
75 0 l . 00 l . 00 0.99 0.99 0.99 0.98 0.98 0.97
400 0.99 0.97 0.95 0.93 0.91 0.89 0.86 0.84
Eq. 22
75 0 l . 00 0.99 0.99 0.98 0.98 0.97 0.97 0.96
544

4.5.Stability in trapezoidal channels

Let us consider a trapezoidal channel(1) with horizontal bed


and side slopes making a angleS to the horizontal.
Simple theoretical considerations indicate that the resis-
tance of side slopes to erosion is lower than the resistance
of the bed, the respective coefficient being given by:

CS = cos S 1 -
tan
2
e
2
tan <I>
(23)

(expression that is known as the Lane formula) and indicate


also that:
d
Ce (24 )

d s being the characteristic dimension of a block on a side


slope in the same stability conditions as a block on the bed
having d as characteristic dimension.
Ulrich, also considering a alternative theoretical model,
proposes instead of Eq. 23:

2
tan
2
(25)
tan

and he likewise proposes in this case the use of the bearing


angle instead of the angle of repose. Moreover Ulrich states
that, if the bearing angle with the nominal value 75 0 is
used, the Eqs.23 and 25 are practically equivalent to:

co = cos S (26)

(1) other cross-section shapes are unlikely in rockfill


channels
545

4.6. Stability in bends

There is not much information on this subject. We think that


this has two causes:

the ratio rib, r being the bend radius at the channel


centerline and b the channel surface width, probably
does not have a very low value in rockfill channels;
thus stability conditions in bends are not very changed
the experimental study, which would be necessary com-
pletely clarifying this question, would be extremely
long.

A United Nations publication of 1973 proposes an increase


of one third in the velocity against blocks. Maynord, 78,
proposes a increasing factor for the block dimension in bends
given by 3.2/Yr/b (on the basis of tests in a channel with
side slopes of 1 to 2 and a bend angle of 60 0 ) .
These two indications lead to the same result when rib is
close to 3.
According to Maynord's indication, the block dimension will
be increased by only 10% when rib is nearly 8.5.

4.7.Effects of lining thickness, gradation, shape and


specific gravity

a) Lining thickness

Several indications on this matter may be synthetized as


follows:

the minimum thickness should be at least d max , 1.5


dSO or 30 cm
dimensions larger than 2 d max are not usual
Maynord et al.on the basis of a limited number of
tests, propose a reduction of d if there is an increase
in thickness in relation to the minimum thickness (1 x
d max ): if the thickness is 1.5 d max d can be reduced by
25% (35% if the thickness is 2 d max ).

b) Gradation

From the stability point of view a well-graded mixture has


an advantage. If is evident, however, that in practice it is
difficult to apply very precise rules (and it is still harder
to prove, in a general way, that theses rules are really nec-
essary) .
546

We can establish the following table to compare ds/d values:

tan 8
1/4 1/3 1/2 1/1. 5
<I>

400 1. 08 1. 15 1. 3 9 1. 98
Eq.23
75 0 1. 03 1. 06 1.13 1. 22
40 0 1.13 1. 25 1. 73 3.26
Eq. 25
75 0 1. 03 1. 06 1. 14 1. 24

Eq. 26 1.03 1. 05 1. 12 1. 20

The coefficients that appear in this table only consider


"the decrease in riprap stability that results from the grav-
ity component acting down the side slope" (Maynord et al.).
These authors list other factors that are favourable to the
stability of side slopes: non-uniform velocity distribution
in a trapezoidal cross-section, stronger interlocking among
blocks on side slopes and adaptation of side slopes to small
local failures. Consideration of a single factor (the first)
therefore leads to pessimistic conclusions that reality does
not confirm. Maynord et al. mention several authors (Ulrich
among them) who have experimentally verified that expressions
such as the Lane formula, employed with <I> = 400, are very
conservative. In fact, Maynord had already experimentally
verified in 1978 that side slopes 1 to 3 (or less) do not im-
ply a decrease in stability in relation to the bed.
547

Several indications from the literature can be synthetized


as follows:

ratio between d max and d50: 2 to 5


d max should be smaller than 16 dIS
Maynord et al. note that "most prototype riprap used
for open channels has ratios of d85/d15 ranging from 2
to 4".
c) Shape

A USA Corps of Engineers publication of 1970 indicates:


angular and cubical shapes are better than rounded and
elongated ones
blocks whose largest dimension is more than three
times greater than the smallest should be avoided
the number of blocks with the largest dimension /
smallest dimension ratio between 2.5 and 3 should be
less than 1/4 of the total number of blocks.

d) Specific gravity

Its influence, as already mentioned, is not great. However,


when s differs notably from 2.65 (and, obviously, an expres-
sion without the parameter s is being used), a corrective
factor with the value 1.65/ (s-l) (1) could be applied to the
characteristic dimension.
For instance, for blocks with specific gravity 2.5, charac-
teristic dimension should be increased by 10% (or nearby 13%
according to Maynord et al.) .

4.8. Conclusions

In the case of bidimensional channels, with horizontal or


quasi-horizontal bed and "normal" relative roughness (up to
0.05), Eqs. 11 (with CI = 0.86 and associated with Eq. 13),
12 (associated with Eq. 13 too), 14, 16 (associated with
Eq.6) and 17 can be used. Fig. 7 enables us to evaluate the
differences that result from using one or other of these
equations.
In the case of high relative roughnesses (E > 0.05) we
think that the Izbash expression (Eq. ~1) and the Olivier ex-
pression (Eq. 19) are applicable, the latter taking into ac-
count the values of f from Fig. 6. Instead of us ing directly
these expressions it is possible to apply the graph of Fig.
8, which provides, according to I zbash and Olivier, the di-
mensional coefficient k (m. sec units) of U = k Yd as a func-
tion of E.

(1) [1.65/(s-1: 1 1.25 according to Maynord et al.


548

Except in the case of very steep channels it does not seem


necessary to reduce the critical velocity. If this is decid-
ed, Section 4.4 explains how to do i d 2) .
d(m)
1. , - - - - - - - - - - - - - - - - - - ,

00

04 -

o.
oL--~---"--~-~--~-~-~~
o o. 10 14
1. cKr:=m''--::=____--'U:.c{:cm'c0::'--'- - - - - - - - ,

00

04

01
oL-~~~~L-~~-L~_L_~-L~_~

o 1 a: 3 4 5 eTa 0 10 11 12 13 '"
d(m) U (m/a)

1.0

0.8 ----~--- ....

o.e ---~---

0.2 - - - - -

oL-~~_L~_~~-L~_L_~~~_~

01234581891011121314
ct{rn} U (m/a)

oL-~~-L~_L_~-L~_~~~~_~

012348 o 7 0 "011121314
U (m/a)
cKnII
1.1 =0.05
1.0 -----

0.8 -

0.0

0.4 ----

0.1

oL-~~_L~L-~

o 23-4887891011121314
U {mI.'

Figure 7. Comparison of stability formulas

(2) no correction of this type should be introduced if


Olivier's expression is used
54')

As regards trapezoidal channels, it does not seem necessary


to increase block dimension on side slopes if they are 1 to 3
(or less). For side slopes 1 to 2 an increase of 15% in block
dimension is proposed (30% in the case of 1 to 1.5). These
increases agree with those proposed by Maynord et al., al-
though they are a little more conservative.

-~

K I I
IH-K) I !

~--
~~
-....;;;:
~ f=:::::::: Ilge~:;~u 10
I--
4
r-- lCCeld;n, Ie
OllVIR
I
1 I

o 10 0.10 0.30

Figure 8. Stability in case of high relative roughness

In Sections 4.6 and 4.7 some indications are given concern-


ing stability in bends and the effects on stability of lining
thickness, gradation, shape and specific gravity.

5.SEEPAGE FLOW

5.1. Preliminary remarks

Seepage in rockfill is hardly ever laminar (if we consider,


as said in Section 1.1, 1 cm as minimum dimension of rockfill
particles) .
It should however be noted that this statement is not true
in the case of the typical modern rockfill embankment, in
which a permeability coefficient higher than 10-3 cm/s
(Maranha das Neves and Veiga Pinto, 87) is enough for the
embankment be considered a rockfill embankment. I. e., this
permeability coefficient limit is relatively low and so sev-
eral rockfill works with occurrence of laminar flow are pos-
sible.
This question is related to the choice of the characteris-
tic dimension adequate for studies on seepage flow. It is
550

typical to mention d with a low subscript, for instance dIS,


dl0 or even d5. In fact, if the situation is one of small
particles filling the voids among far larger particles, it is
obvious that the controlling particle size concerns small
particles. But, as already said, the situation that we anal-
yse is a different one. Leps, for instance, referring to the
studies of Wilkins, 56 on high permeability media, proposes
the choice of dSO.
It is typical to classify laminar flow and turbulent flow
on the basis of the dimensionless parameter Reynolds number
whose classical definition is:

U 0
Re (27 )
1)

1)being the kinematic viscos ity of water.


The Reynolds number is a kind of measure of the influence
of viscosity forces on flow, its value decreasing when this
influence increases.
Between rough turbulent flow (i proportional to U2 and lam-
inar flow (i proportional to U) there is a transition zone.
Some authors (for instance Kovacs, 81) refer to two transi-
tion zones, even in the case of seepage flow, by analogy with
non-seepage flow (one of them corresponding to the unstable
transition from laminar flow to non--rough turbulent flow
and the other to the transition non-rough to rough turbulent
flow). In our opinion it is not necessary, when analysing
seepage flow, to consider two transition zones.In fact, if we
examine the schematic Moody diagram on Fig. 9, we can verify
that the second of these zones, in the case of high friction
head losses (i.e., the present case), shows a small develop-
ment, and it is possible to associate this zone with the
first one.
On the other hand the instability of the first zone should
not occur in flow through porous media: as Oliveira Junior,
83, notes, the irregularity of the porous medium implies
that in the same volume of the medium zones with turbulent
flow and zones with laminar flow can co-exist. Thus, the
mildness of the transition.
Reynolds numbers that characterize the transition zone are
much lower than the corresponding ones for non-seepage flow
(say, the usual range 2 000 - 10 000 is reduced by at least
one order of magnitude) (1) .

(1) more on this matter can be found in Section 5.4


55!

It is possible to list three reasons for this:


- the boundary irregularity is propitious to the develop-
ment of t urbulence(1)
- occurrence of non-linear laminar flow (see, for in-
stance, Scheidegger) due to local head losses that ex-
ist in flow through porous media
in flow through porous media, unlike non-seepage flow,
the effective distance travelled by the flow may in-
crease when head increases (see Annex III).

O.071-------t~?t=.........-+_-------------------~O.05

Figure 9. Schematic Moody diagram

The homogeneity and isotropy hypothesis(2), already .men-


tioned in Section 2.4, is considered valid for seepage flow.
Besides this hypothesis, it is also considered that the medi-
um is saturated. As we know, any of these circumstances may
not occur in a rockfill work.

(1) it should be noted that some experiments have shown the


"symmet-rical" phenomenon: flows from which causes of
perturbation are removed remain laminar with high Reynolds
numbers
(2) i.e., not only has e the same value at any point, but
also at any point permeability is the same in all directions
552

5.2. Mean hydraulic radius of the voids and mean ve-


locity in the voids

There are several models for studying the permeability of


porous media (see Scheidegger). One of them, which seems to
be adequate in the field of civil engineering, is based on
the concept of mean hydraulic radius of the voids, m, defined
as the ratio of volume to surface of the void space(1 )
From this definition, supposing irrelevant the area of the
contact zones among blocks, we have:
m = e (28 )

Void ratio, as said in Section 2.4, depends on block shape,


gradation and degree of compactness. Let us suppose that we
can estimate the void ratio corresponding to each association
of these three parameters, i.e., that e is a datum. Thus,
knowledge of m implies knowledge of sv.
The experimental determination of Sv is not practical. The
method is time-consuming, not very precise and the sample
should not be too small (sometimes 100 is considered a ade-
quate sample dimension in this type of studies) .
As mentioned (Section 2.4) Sv depends on block shape and
gradation and is inversely proportional to d.
Thus we have:
ed (29)
m
c'

in which C' is a dimensionles coefficient depending on block


shape and gradation.
The main interest of this expression is to make it evident
that the void ratio must appear associated with d in the laws
of seepage flow. An expression of the type m = d/8 (suggested
by Leps) is an oversimplification.
It should be stressed that the introduction of the parame-
ter e does not cause any supplementary difficulty. In fact,
we usually want to know the seepage discharge Qs and, in or-
der to know it, it is necessary to know the void ratio (or
the porosity). When studying seepage flow in rockfill it is
more logical to use Uv , mean velocity in the voids, than V
(seepage velocity or bulk velocity) .
Thus we have:
(3 0)
Qs = Sa V = n Sa Uv
Sa being the cross-sectional area (apparent or total) .

(1) it is easy to prove that another concept, the concept of


intrinsic permeability, is equivalent; in fact, intrinsic
permeability is a hydraulic radius squared
553

It should be noted that this expression is based on:

uv = V (31 )
n

Eq. 31 is commonly used and seems evident. It results, how-


ever, from an assumption related to the homogeneity and
isotropy hypothesis, known as the Dupuit-Forchheimer assump-
tion (see Scheidegger). It is opportune to quote Shapiro, 87:
"the equivalent porous medium conceptualization assumes that
one continuum velocity is representative of the average of
all velocities in the void space of the medium".
From the above considerations we can conclude that it is
decisive for the study of rockfill seepage flow to have in-
formation about void ratios.
Considering again Eq. 29, we can observe that this expres-
sion would improve considerably if it assumed the form m = f
(e, d, Cs , Cu ), admitting that a single shape coefficient and
a certain Cu (uniformity coefficient) characterize, respec-
tively, shape and uniformity.
Difficulties analogous to those we face when determining
specific surface appear when determining shape coefficients.
On the other hand, as we will see, uniformity does not have
in general a great influence on rockfill seepage flow.
Thus, it seems that there is not any better expression than
a expression of the type of Eq. 29. Linford and Saunders, 67,
indicate C' = 8.3 for angular particles (almost uniform). De-
terminations made at LNEC (small samples) are in agreement
with this value and point to a value between 6.0 and 6.5 for
uniform rounded particles.

5.3. Turbulent seepage flow

There are several expressions concerning this type of flow.


Scheidegger, Bear, Kovacs and Oliveira Junior present some of
them. It should however be observed that these expressions,
in general, have been obtained by using regular-shaped parti-
cles and in contexts that are different from that of civil
engineering.
In this last context we can mention, among others, the ex-
pressions of Wilkins, of Tepaks(1) (both presented in 1956)
and of Parkin, 62 (2). As far as we know, the Wilkins expres-
sion is the one most commonly used in practice.
Tests concerning this type of flow have been carried out at
LNEC (Martins, 90). The main information on these tests can

(1) mentioned by Isachenko and Lyubimova, 73


(2) and, more recently, the expression proposed by Jain et
al., 89
554

be found in Annex IV but we think that it is convenient to


present here the following points:
the particles that simulated rockfill were irregular-
shaped, angular or rounded, uniform or non-uniform,
with relatively large dimensions (approximately 1 to 13
cm)
the experimental installation was also relativply large
(internal diameter of the pipe containing the par-
ticles: 50 cm).

Eq. 27 is equivalent to:

4U v m
Re (32)
U

In the LNEC tests the minumum Reynolds number was close to


300. No deviat ion in relation to a quadratic law was ob-
served. This agrees approximate ly either with the Kovacs
graphs (which syntethize nearly all known results on this
type of flow) or with Gordienko statement (quoted by
Isachenko and Lyubimova) about the applicability of a
quadratic law if, simultaneously, d > 1 cm and i > 0.1 (and,
also, with the values of Re mentioned by Bear). We can add
that, as we will see, it is not important, in practice, to
define precisely the value of this limit.
Taking the above into account, the resistance law of turbu-
lent seepage flow assumes a form similar to that of the Dar-
cy-Weisbach expression (Eq. 5)

1
i = fs (33)
4m

in which fs is a dimensionless coefficient that is constant


if we accept the already mentioned hypothesis of the surface
roughness of the blocks not being relevant.
Combining Eqs. 29 and 33 we have:

Uv = ~ 4
f sc 1
Y2 g e di (34)

that we can transform in:

(35)
U v = C f (C u) Y2 g e di

in which C is a dimensionless coefficient that depends only


on block shape and f(C u ) a function that depends only on uni-
formity degree, which becomes unity for uniform materials.
555

As already mentioned, it is not possible, at present, to


express C as a function of Cs . On the basis of LNEC tests we
determined C = 0.56 for angular particles (C s ~ 0.7) and C
= 0.75 for rounded particles (C s ~ 0,95). The ratio 0.75/0.56
is close to the corresponding ratio 1.4 mentioned by Wilkins.
If we compare values of Cs with values of the coefficient
C, it seems that the change in the velocity through voids is
due not only to shape roughness - rounded blocks with smaller
shape roughness than angular blocks - but also to surface
roughness rounded blocks likewise with smaller surface
roughness. As far as we know, no experimental study has dealt
with this subject.
On the basis of only three sets of particles (each with a
different value of Cu ) and considering adequate the func-
ex
f (C ) l/Cu with C
tion u u

1
0.26
(36 )
Cu

The exponent 0.26 needs more experimental confirmation.


However, in rockfill without fines, uniformity degree does
not seem decisive, since C u does not assume high values (we
should remember the Maynord et al. observation, quoted in
Section 4.7, about it not being common for dS5/d15 to exceed
4 in rockfill works subject to water flow).
Furthermore we can observe that the sample a) (Annex I)
presents a low Cu although with blocks of very different
sizes. This sample is not unrealistic. In fact Thompson and
Campbell have worked in the field with rockfill materials
whose representative samples were of this type.
In Annnex V LNEC results and Wilkins results are compared
(Jain et al. results too)

5.4. Transition zone

Even in rockfill defined as mentioned (minimum dimension 1


em) non-quadratic flow may occur (obviously, the same will
happen with particles of, say, some mm) .
Limits of the transition zone in function of Re have al-
ready been mentioned in Section 5.1. It should be noted that
these limits are usually given in function of a Reynolds num-
ber defined not as in Eq. 32 but according to:

Vd
Re (37 )
u
556

In fact, taking into account the usual values of nand C ' ,


the two definitions lead to similar values, at least from the
point of view of establishing the transition zone limits that
vary appreciably (see, for instance, Scheidegger).
In rockfill hydraulics this way of expressing the above
limits does not seem the most convenient. In fact, Re depends
on velocity and velocity is calculated in different ways ac-
cording to the type of flow (laminar or turbulent or of tran-
sition) .
It is easy to prove that it is possible to modify the Moody
diagram in order to obtain a more adequate diagram for the
study of seepage flow. This diagram is schematically repre-
sented in Fig. 10.

r Darcy law

5

4

..
'" .. -
c mm
o 2.3
o 6.0 mm
c
3
e
I II
c 011 mm

~
~

2 !\. :
...
COo Cl 00
v ~

o ... 1 2 5 6 7 8 9 l(
ci
Xo

Figure 10. Modified Moody diagram


557

It is a non-logarithmic diagram. The parameter ~ is given


by:
Uv quad
~= (38)
Uv tran

where Uvquad is the value of Uv according to a quadratic law


and Uvtran is the value of Uv in the transition zone. The di-
mensional parameter X (with d in mm) is given by:

X = e d i 1/3 (39)

It should be noted that we can estimate the hydraulic gra-


dient i on the basis of the known conditions of head and dis-
tance (see Annex II)
With the purpose of studying the transition zone, tests
were carried out at LNEC, in the installation mentioned, us-
ing smaller angular particles, uniform of fairly uniform. The
experimental points are included in Fig. 10 as well as the
approximate limits of the transition zone in terms of the pa-
rameter X. It was verified that the coefficient C of Eq. 35
has a tendency to decrease when d decreases (Wilkins pointed
out the same). Tests showed that the value 0.42 is adequate
to C in this case.

5.5. Conclusions

Eqs.35 and 36, with C = 0.56 for angular particles and C =


0.75 for rounded particles, make it possible to calculate
turbulent seepage flow.
Laminar, transition and turbulent zones are demarcated in
Fig. 10. Limit values are nominal values.
Using Fig. 10 associated with Eq. 35 (with C O . 42) i t is
possible to estimate Uv in the transition zone.
Eq. 36, probably less significant for practical use, needs
more experimental confirmation. More experimental work is
also necessary as regards the influence of particle shape and
an eventual influence of surface roughness.

ACKNOWLEDGEMENTS

We wish to thank the following persons for their support:

Maranha das Neves, Veiga Pinto and Delgado Rodrigues of


the Geotechnique Department (LNEC)
55R

Francisco Palma, Esmeralda Gonzalez, Heleno Cardoso,


Mendes de Carvalho and Maria Jose Cavaca of the Depart-
ment of Hydraulics (LNEC)

George Dykes.

REFERENCES

Bathurst, J. C. - Flow resistance of large-scale roughness -


Journal of the Hydraulics Division no. 12
(December 1978)

Bear, J. Dynamics of fluids in porous media - McGraw-Hill


Book Company - New York, 1975

Chow, V.T. - Open-channel hydraulics - McGraw-Hill Book Com-


pany - Tokyo, 1959

Corps of Engineers - Hydraulic design criteria - n.d.

Corps of Engineers - Hydraulic design of flood control chan-


nels - 1970

Gordienko, P.I. - The influence of channel roughness and flow


states on hydraulic resistances of turbulent flow -
Journal of Hydraulic Research no. 4, 1967

Graf, W.H.- Flow resistance over a gravel bed; its conse-


quence on initial sediment movement in "International
workshop on fluvial hydraulics of mountain regions" -
I.A.H.R. - Trent, 1989

Henderson, F.M. - Open channel flow - The Macmillan Company -


New York, 1966

Isachenko, N.B. and Lyubimova, V.D. - Estabilidade dos blocos


no paramento de jusante de uma barragem de enrocamento
sob a ac<;:ao de infiltra<;:6es em regime turbulento (in
Portuguese) - Translation no. 515 - LNEC, 1973

Izbash, S.V., and Khaldre K.Y. - Hydraulics of river channel


closure - Butterworths - London, 1959

Jain, S.C., Holly, F.M. and Lee, T.H. Head loss through
porous dikes - University of Iowa - Iowa City, 1988
559

Kovacs,G. - Seepage hydraulics - Elsevier Scientific Publish-


ing Company - Amsterdam, 1981

Leps, T.M. - Flow through rockfi1l in "Embankment dam engi-


neering" - John Wiley and Sons - New York, 1973

Linford, A. and Saunders, D. - A hydraulic investigation of


through and overflow rockfill dams - BHRA - 1967

Maranha das Neves, E. and Veiga Pinto, A. - Enrocamentos. Ac-


tualizac;::ao de conhecimentos, estudos experimentais e
aplicac;::ao em barragens e vias de comunicac;::ao (in Por-
tuguese) - LNEC - Lisbon, 1977

Maranha das Neves, E. and Veiga Pinto, A. - Enrocamento com-


pactado: um novo mater ia 1 de construc;::ao (in Por-
tuguese) - LNEC - Lisbon, 1987

Martins, R. - Turbulent seepage flow through rockfill struc-


tures - Water Power and Dam Construction no. 3, 1990

Maynord, S.T.- Practical riprap design - Waterways Experiment


Station - Vicksburg, 1978

Maynord, S.T., Ruff, J.F. and Abt, S.R. - Riprap design -


Journal of Hydraulic Engineering no. 7 (July 1989)

Neill, C.R.- Guide to bridge hydraulics University of


Toronto Press - Toronto, 1973

Oliveira Junior, A.C. Escoamento turbulento em meios


porosos (in
Portuguese) - Escola Politecnica da Universidade de S.
Paulo - S. Paulo, 1983

Parkin, A.K.- Rockfill dams with inbuilt spillways. Hydraulic


characteristics - Water Research Foundation of Aus-
tralia - Melbourne, 1962

Parkin,A.K.- Rockfill dams with inbuilt spillways. Stability


characteristics - Water Research Foundation of Aus-
tralia - Melbourne, 1963

Parkin, A.K., Trollope, D.H. and Lawson, J.D. Rockfill


structures subject to water flow - Journal of the Soil
Mechanics and Foundations Division no. 6, 1966

Pereira - J.J.C.P. - Caracterizac;::ao geotecnica de macic;::os ro-


chosos (in Portuguese) - Universidade Nova de Lisboa -
Lisbon, 1985
560

Quintela, A. C. Hidraulica (in Portuguese) Funda<;:ao


Calouste Gulbenkian - Lisbon, 1981

Scheidegger, A.E. - The physics of flow through porous media


- University of Toronto Press - Toronto, 1960

Scott, R.F.- Principles of soil mechanics - Addison-Wesley


Publishing Company - Reading, 1963

Shapiro, A.M. - Transport equations for fractured porous me-


dia in "Advances in transport phenomena in porous me-
dia" - Martinus Nijhoff Publishers - Dordrecht, 1987

Stephenson, D. - Rockfill in hydraulic engineering - Elsevier


Scientific Publishing Company - Amsterdam, 1979

Thomas, H.H. - The engineering of large dams (chapter 15)


John Wiley and Sons - London, 1976

Thompson, S.M. and Campbell, P.L.- Hydraulics of a large


channel paved with boulders - Journal of Hydraulic Re-
search no. 4, 1979

Ulrich, T. - Stability of rock protection on slopes - Journal


of Hydraulic Enginering no. 7 (July 1987)

United Nations - Design of low-head hydraulic structures -


New York, 1973

Veiga Pinto, A. - Caracteristicas de resistencia e deforma-


bilidade dos materiais de enrocamento (in Portuguese)
- Revista da Sociedade Portuguesa de Geotecnia no. 27
(November 1979)

Wang, S. and Shen, H. W. Incipient sediment motion and


riprap design - Journal of Hydraulic Engineering no. 3
(March 1985)

Wilkins, J. - Flow of water through rockfill and its applica-


tion to the design of dams - Proceedings of the Second
Australia - New Zealand Conference on Soil Mechanics,
1956

Zharkov, Y.G. - Determination of noneroding velocity with a


given reliability for noncohesive coarse-granular
soils - Hydrotechnical Construction no. 6 (June 1985)
561

NOTATION

A - particle area projected on a plane normal to the


flow
b - channel surface width
C - dimensionless coefficient (law of turbulent
seepage flow)
C - dimensionless coefficient (relating m to ed)
CD - drag coefficient
CI - Izbash coefficient
Cs - shape coefficient
Cu - uniformity coefficient (usually d 60/dlO)
Ca. dimensionless coefficient (critical velocity in
the case of positive slope)
Ce - dimensionless coefficient (block dimension on a
side slope)
d - characteristic dimension of a particle
D - hydraulic diameter
-
d - mean diameter (of a single particle)
dl,d2,d3 - dimensions of a particle measured according to
three orthogonal directions
dSO, etc. - diameters from the grain-size distribution curve
de - equivalent diameter
dm - mean diameter (of a set of particles)
d max - maximum dimension of blocks
dn - nominal diameter
ds - characteristic dimension of a block on a side
slope
e - void ratio
f - friction factor (Darcy-Weisbach factor)
FD - drag force
fs - friction factor for seepage flow
562

g - acceleration due to gravity (considered 9.8m/s2)


h - flow depth
i hydraulic gradient
k number of particles in a given set; dimensional
coefficient (rockfill stability with high rela-
tive roughnesses)
k' - absolute roughness
Ks - Gauckler-Strickler resistance coefficient
m - mean hydraulic radius of the voids
n - porosity
q - discharge per unit width (critical conditions)
Q - discharge
Qs - seepage discharge
r - bend radius at channel centerline
R - hydraulic radius
Re - Reynolds number
s - specific gravity of rocks
S - cross-sectional area of flow (effective)
Sa - cross-sectional area (apparent or total)
Sv - specific surface (surface area per unit volume)
U - mean velocity; mean velocity in critical condi-
tions
Ub - mean velocity in the vecinity of the block
(critical conditions)
Uv - mean velocity in the voids
Uvquad - seepage velocity according to the quadratic law
Uvtran - seepage velocity in the transition zone
Ua - mean velocity in a channel with bed slope tan a
(critical conditions)
V - volume of a particle; seepage velocity (or bulk
velocity)
X - parameter in the modified Moody diagram
563

a - bed slope angle to the horizontal (channel or


downstream slope of a overflow dam); exponent in
the function f (C u )
E - relative roughness

~ parameter in the modified Moody diagram


y unit weight of water
U - kinematic viscosity of water (considered 1.15xlO- 6
m 2 / s)
'to shear stress at solid boundary
<\> friction angle
8 side slope angle to the horizontal
564

ANNEX I

EXAMPLES OF CALCULATING THE CHARACTERISTIC

DIMENSIONS OF BLOCK SETS

sieves: 0.15 m 0,25 m 0,35 m, etc.

sample a)

11 blocks with the following values of d(m) :

0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00
1.10

arithmetic mean of diameters 0.60


median diameter (or Wolman diameter) (1 ) 0.60
dm 0.73
d 10 0.65
d 50 0.97
d 60 1. 01
d 75 1. 07
d 90 1.12
d 60 /d 10 1. 55

sample b)

5 blocks:

0.40 0.50 0.60 0.70 0.80

arithmetic mean of diameters 0.60


median diameter (or Wolman diameter) 0.60
dm 0.63
d 10 0.50
d 50 0.72
d 60 0.75
d 75 0.79
d 90 0.83
d 6 Old 1 0 1. 50

(1) according to Thompson and Campbell, 79


565

ANNEX II

HYDRAULIC GRADIENT

Hydraulic gradient is defined as the unit piezometric head


loss of a flow. Representing the pressure(1) variation (in
terms of water column, according to: pressure unit weight
of water x height of water column) in the unit of length, the
hydraulic gradient is dimensionless.
The piezometric head is equal to pressure/ unit weight plus
a vertical height above datum.
The hydraulic grade line (hydraulic gradient line) of a
flow is drawn at a heigh above datum equal to the piezometric
head and can be indicated by the readings on piezometers
along the flow.
The energy line (total head line)
is drawn at a height
above the hydraulic grade line equal to U2 /2g(2) (velocity
head) .
Let us consider an uniform open channel flow. In this flow
the energy line, the hydraulic grade line (coincident with
the free surface) and the bottom profile are parallel lines.
Thus, there is no ambiguity in utilizing a single symbol,
i, for the hydraulic gradient or for the head loss per unit
length, as we do in this chapter.
It should be noted, however, that strictly speaking,it is
not i = tan a but i = sin a (since, as is logical, the
pressure variation is referred to the path line and not to
its projection on a horizontal plane) .
Usually this point has no practical interest, since tana
only differs from sina more than 5% for bed slopes larger

(1) relative pressure, in the context of this chapter


(2) not considering a coefficient generally close to unity
566

than approximately 32%.

energy line
hydraulic orade line and tree surface

--'-----_~ channel bottom

Furthermore it should be noted that for low velocities,


such as seepage flow velocities, energy line and hydraulic
grade line are practically coincident.
Finally it should be noted that many flows, among them many
seepage flows, can be schematized according to:

Rldraur
Ie grade
line }H
,l L ~

In these conditions the hydraulic gradient can be given by


i1H/L.
567

ANNEX III

A REASON FOR APPARENT NON-LINEARITY IN


LAMINAR SEEPAGE FLOW (CONCEPTUAL EXAMPLE)

Let us consider a one meter long tube filled with a particu-


lated medium in which a laminar seepage flow occurs. This
flow is caused by a constant head difference, ~H = 1 m.ln
these conditions i = 1.
We consider that, as a result of the particulated medium
characteristics, Uv = 1 mm/s.
We can write the laminar flow law:

i K Uv

or i Uv (U v in mm/s)

Let us suppose that ~H becomes 2 m (flow remaining lami-


nar). If the above equation were valid, without any change,
we would measure Uv = 2 mm/s.
In fact, the effective length may increase when Uv increas-
es, because new paths of higher tortuosity among the parti-
cles may arise. Let us suppose, continuing this conceptual
example, a length increase of 10% .
The effective value of i becomes not 2/1 but 2/1.1 = 1.82.
Thus, the resulting value of Uv would be not 2 mmls but 1.82
mm/s and the output of this conceptual test would be:

test output

effective apparent i Uv (mm/s)

1 1 1

1.82 2 1.82

This output would be explained not by i Uv but by i


1.16, although the flow is still laminar.
568

ANNEX IV

TESTS FOR THE QUADRATIC ZONE

During these tests, use was made of:

eight sets of uniform angular particles (crushed mate-


rial)
three sets of uniform rounded particles (alluvial grav-
ell
three sets of non-uniform angular particles (with Cu =
2, Cu = 4 and Cu = 8, each with the same characteris-
tic dimension: 63,S mm),

The apparent volume of each of the sets was 0,34 m3 approx-


imately, The characteristic dimension of uniform materials
was defined by d (d n + d n -l)/2,where d n and d n -l are two
consecutive sieve diameters, d was varied between 11 and 127
mm,
The photo shows part of the experimental facility which was
used for the study:

--the internal diameter of the smooth pipe containing the


material is 500 mm
sections A and E are 1,75 m apart and confine the rock
fragments
the measuring reach (1 m long) is situated between sec-
tions Band D; eight pressure taps were placed in each
of these sections, and visual readings of the corre-
sponding piezometer tubes permit the calculation of i
pressure taps were also placed, for control, in each of
the sections A, C and E,
569

It was verified (by preliminary testing) that the particle


arrangement, keeping e constant, does not greatly modify the
discharge.

The range of the parameters has been:

Uv from 0.061 to 0.909 m/s


e u from 1 to 8
e from 0.551 to 0.925
i from 0.096 to 1.420(1)
Re from, approximately, 300 to 50 000

Values of Uv measured and calculated by Eq. 35 are repre-


sented in the figure.

Uv
measured
[m Is)
1.00

!

0.80
0;,
0
~o

0.60 :"
~
"0
..........
,,:n..
II'

.:t~
0.40 MO~4
_ cl!'ii
angular materials

_,s.""o
.:..~~"

i$f
o rounded materials
f.~;' non-uniform materials
0.10

0.00 0.00 [20 0.40 0.60 0.80 1.00 Uv calculated


[m/s)

(1) according to Parkin et al., 66, "energy gradients in


rockfill structures will generally be in the range 0.1 to
1.0"
570

ANNEX V

COMPARISON OF RESULTS FROM EQ. 35 AND


FROM THE EXPRESSIONS OF WILKINS AND JAIN ET AL.

Let us consider angular and uniform particles.


In these conditions the Wilkins expression can be written:

Uv = 1.18 Y2 g m i 0.54

In the same conditions Eq. 35 is:

Uv 0.56Y2gedi

If we introduce in this expression m = ed/8.3 (according to


Linford and Saunders, 67) we have:

Uv = 1. 61 h g m i 0.5

We therefore conclude that Eq. 35 gives Uv values higher


that the Wilkins expression (increases varying from 50% to
36% when i varies from 0.1 to 1) .
In fact the increases are lower: Wilkins carried out addi-
tional tests, with larger particles, suggesting in this case
an increase close to 28% for the coeficient of his formula.
The expression proposed by Jain et al., 88, obtained with
angular and uniform particles of large dimension (close to 30
cm), constituting a dike, may assume the form:

Uv = 0.54 Y2 g e d i 0.58

Eq. 35 also gives Uv values higher that Jain et al. expres-


sion (increases varying from 25 % to 4% when i varies from
0.1 to 1).
CHAPTER 18
THROUGH AND OVERFLOW ROCKFILL DAMS
A. K. PARKIN

I. I NTRODUCTION

The o ri gin~ of through and overflow rockfill dams can be traced, in Australia at least, 10 the
Cascade Dam (Fig. I). built 10 supply sluicing water to the Bri seis alluvial tin mine at Derby
in north-cast Tasmania (Anon, 1989). It was to enter the annals of history as the only large
dam in Australia to fail causing loss of life. However, despite this disastrous beginning, the
ensuing sixty years have seen the techniques of mesh protection develop 10 the siage where
flood flows can now be passed over rockfil1s with confidence. and where such provision is
almost a normal feature in the construction of rockfill dams. These techniques now permit
substantia! economics \0 be made in the scale of river diversion works, and even the total
elimination of such works in some cases.
Cascade Dam was, in principle, a concrete-faced rockliU (eFR) dam, constructed to
dimensions as in Fig. 2, but with a 70 m central gap in the 0.5 m high parapet wall to serve
as a spillway. This discharged directly onto thc downstream face, with an estimated capacity
of 200 cumecs, 1 generated from a catchment of some 30 km 2 In order to accept these
overflows, the main body of the dam was constructed of granitic rocklill, containing boulders
up to 2 or 3 m1 in sizc, placed to a void ratio of about 0.46. Evidently this was largely hand-
placed in order to achieve the steep side slopes.
On 4th April 1929, the dam was overtopped and failed during an extreme rainfall event,
causing the loss of 14 lives, at which time the spillway flow was estimated to be 6I.X.l eumces l .
Apart from the inadequatel y estimatcd design flood. the steep unprotected downstream slope
ensured that the structure would have a bleak future. Subsequently in 1934 the dam was
reconstructed. with the downstream face flattened to 1.75 : I and additional spillway capacity
(McKeown. 1938). It continues to function, and is again notewonhy in being one of the few
failed dams of such size ever to be rebuilt.

I. At the design flow, il was evidemly imended that the parapet wall would be totally
submerged by 0.5 m or so. The estimated peale flood flow may n01 be reliable, or may
have been a lfaTIsiem surge as discusscd by Bamen (1989).

51'
E. Mur{milu das Ne'es (rd.l. Ad,"UI/(."u ill Rndjill SI""( llIre~. 571-592.
CI 199] K"mer Acadr",j, Publishers.
572

Fig. 1. Locality plan, Tasmania, Australia

2. Early Developments in Flood-protected Rockfills

Experiencc gained in the 1930's with wire gabions in river training works inspired the first
reported use of bar-reinforced rockfills during the construction of three embankment dams in
Mexic0 2 , as rcported by Wciss (1951). Of these, San Ildefonso Dam (1939), a CFR dam
protectcd by a rather open grid of 19mm bars, was overtopped at a height of 11 m by a flow
of some 200 cumecs with insignificant effect, but after the bars were recovered for use in the
facing a further flood washed out 7000 m 3 of rock. EI Azucar Dam (1941) was also
overtopped, suffering some damage due to an inadequate height of protection, while Palmi to
Dam (1941) withstood a massive 1400 cumec flood.
A fourth dam (Valsequillo, 1943) falls into a rather different category, being a central core
rockfill in an exceptionally narrow canyon, where a perforated concrete slab was judged to be
the best means of protection against the concentrated flow and flood debris. Of three
overtoppings, the last was the most severe because of the bank then standing at 39m. Internal
pore-pressure lifted and cracked the slab, but caused no failure because of its anchorage into
the canyon walls. This form of protection, however, has not been used again and is unlikely
to be appropriate elsewhere.
Thcse successes in the passage of flood flows over rockfill during construction aroused
immediate interest in the USA where such techniques had been beyond contemplation since
the Johnstown (Pa) disaster of 1889 (Mattes, 1951). However, it would appear that mesh

2. A recent report indicates that the first usc of reinforced rockfill was on two darns in South Africa
incorporating wire gabions (Prins River 1918, and Bellair 1920), both of which were overtopped
several times (ICOLD, 1986).
573

reinforcement still did not find its way into US practice until after the failure of Hell Hole Dam
(California) in 1964, in which overtopping during a 1050 eumee event caused the loss of over
500000 m3 of rockfill (Johnson, 1971; Lcps, 1973)3.

Crest length 137m


Volume 28400m 3

24m
Granite 1: 1
Rockfill
. ().

o o~.o~o
0.' 0 QO
. 0

Granite

Fig. 2. Approximate profile, Cascade Dam

3. The Self-spillway (throughflow) Dam

From the observation that even unreinforced rockfills4 have a useful, if modest, capacity to
pass water, the concept of a self-spillway dam was developed by Wilkins (1956)
for application to a small (l2m high) structure in central Tasmania, for which the design flood
was around 30 eumecs (Fig. 3). As compared with the unfortunate Cascade Dam, this strueture
had an internal spillway, buried under some 2.5m of rockfill and located so as to ensure that
the phreatic surface would not daylight on the downstream slope. In addition to a flatter slope
at the natural angle of repose, a stabilising berm of large rocks was placed at the point of exit,
and the structure continues to function satisfactorily since completion in 1957.
The main rockfill consists of tipped dolerite, which, although uncompacted, is relatively rigid
(settling - Imm/year after 7 years). This is not unusual for structures of this size (Wilkins,
1989), but the use of an inclined clay core clearly offers advantages over concrete in the event
of deformations occurring. The clay core is capped with a concrete sill to form a spillway over
a central section of some 55m, from which water falls freely into the supporting rockfill (which
it does continuously, as this structure has no outlet works beyond an emergency valve).
3. It would appear that the Pit 7 Afterbay Dam (1965) is still the only recorded use of reinforced
rockfill in USA (I COLD, 1986), together with a similar structure for the Yuba River Project
(Sarkaria and Dworsky, 19(8).

4. In this case, rockfill embankments supporting concrete flumes.


574

Fig. 3. Laughing Jack Marsh Dam (Hydro-Electric Commission, Tasmania) (Wilkins, 1956)

On the assumption of a mean panicle size - 300mm and a void ratio of 0.85, critical depth
over a buried weir at the design flow was determined to be

Yc = ~q2/n2g - 0.5 m (1)

wherein q = discharge/unit width and n = porosity), for which the corresponding void velocity
was 2.2 m/sec. A drawdown profile was then calculated, based on a non-linear head-loss
equation from permeameter tests, to reconcile Yc with the assumed reservoir level (F.S.L. +
I.5m) from flood routing studies, with adjustments to geometry and stone size being made as
necessary. However, the applicability of the critical depth criterion in a frictional environment
is open to question, as discussed later.
While the self-spillway concept generally has a restricted application to small flow situations,
it has been incorporated into a proposed tailings dam for the Faro Abandonment Scheme
(Yukon, Canada), which, if built, would be by far the world's largest strueture of this type.
This structure, 50m high, would be required to pass some 530 cumecs over an internal erest
under arctic conditions, where it is required to be maintenance free indefinitely (J.K. Wilkins,
priv. comm.).

4. Laboratory Studies

The functioning of a self-spillway dam has been further examined in a series of model studies,
which served, in the first instance, to define internal regions of flow as in Fig. 4 (Lawson,
Trollope and Parkin, 1962). Energy dissipation occurs through the crest, freefall and
downstream regions, allowing a benign flow to exit from the toe, thus eliminating the need for
energy dissipation works. The design problems include the prediction of water surface profiles
(which determine flood capacity and the distribution ofpore-prcssure and vclocity), the stability
of the downstream toe against erosion, and stability against internal slip failure.
575

Approach

In -built
spillway

Impervious wall

Fig. 4. Regions of Flow in a Self-Spillway Dam (Lawson et aI., 1962)

4.1 Equations of Flow

The Darcy Equation of seepage flow ceases to apply for particles larger than a few mm (as
demonstrated in an earl y paper by Bakhmeteff and Feodorov, 1937), so that a turbulent regime
persists from prototype scale right down to model scale, where aggregates of 10 to 20mm have
been used. In a series of permeameter tests on crushed dolerite in the range 19 to 40mm,
Wilkins (1956) established the head loss equation (m. sec units).

vv - 5 3 mO.s iO.54 (2)

where Vv (= v/n) = velocity through voids5, m = hydraulic mean radius (taken to be d/1O in
this case) and i = hydraulic gradient (for which the exponent was constant).
Although there have been suggestions that Reynolds Number scale effects persist through this
size range6 , Henderson (1956) argues that surface drag is insignificant and that, as a
consequence, it is inappropriate to define m in terms of surface area per unit volume
(suggesting it should reflect only void size). It also follows that Equation (2) is unlikely to
vary with temperature. Energy loss has also been expressed in terms of the Missbach Equation
(e.g. Lawson, Trollope and Parkin), which is essentially an inversion of Equation (2)

i - av G (3)

wherein a = 1.85 and the coefficient a is a function of m and the porosity n; and in terms of
the Forchheimer Equation

5. Some (e.g. Bakhmeteff and Feodorov) suggest this factor should be in n2i3 , but despite the
theoretical arguments, Stephenson (1979) claims that vln still gives the most realistic estimate
of void velocity.

6. Sandie (1961) reponed transitional flow in a IOrnm aggregate (as compared with 19mm
aggregate). However. Stephenson (1979) claims that transitional flow persists up to Reynolds
Numbers of around 10". at which point the exponent of i in Eq. (2) becomes 0.5.
576

i - av + bv 2 (4)

Thc latter has attractions in its rational format, wherein the first term can be associated with
viscous losses and the second term with form drag, which becomes dominant in the turbulent
range. However, the simplicity of a onc-term expression can have computational benefits.

4.2 Hydraulic Control Points

The geometry of flow through a self-spillway dam will, in most cases, be fixed by two
independent hydraulic control points, at the crest and at thc downstream exit (Fig. 4).
Separating them is a freefall zone at atmospheric pressure, wherein i = 1 and whercin latcral
dispcrsion and aeration takes place. At high rates of discharge, the frccfall zone may be
eliminated, at which stage the control points are no longer independent. Howevcr, the
structure is thcn likely to bc in a precarious state with rcspcct LO stability, so that this is not
normally a design consideration.
In the crest region, Wilkins (1956) estimated the crest depth he by means of the
critical depth formula, as normally applied to frictionless flow over weirs, modified
only to allow for the volume of stones (Equation 1). However, this approach was
considered to be inappropriate in a frictional environment, such that the computed
crest velocity was likely to be otpimistie (Henderson, 1956).
Subsequent model studies (Sandie, 1961) have indicated that crest flow is
characterised by the attainment of a terminal velocity, associated with a particular
value of hydraulic gradient in the frictional environment. This gradient can be
taken to be 0.8 (rather less than the maximum of 1 in free-surface flow), in which
case the associated value of he (assuming m = d/10) can be compared with Ye from
Equation 1 in an analysis following Lawson, Trollope and Parkin
3 2 (5)
Ye - he . d/6

It follows that thc crest velocity will be less than that from a critical depth analysis for all
depths of flow greater than d/6. It also follows that discharge will vary linearly with head over
the crest, rather than head to the power 1.5, as confirmed experimentally by Sandie. 7
At the point of exit from the downstream LOe, the exit depth he has also been found, from
experiment, to be proponional LO discharge, provided that it exceeds the depth at tailwater. It
not, then tailwater level will provide the downstream control.
In the event that he > htw (tailwater depth), the phreatic line will exit tangentially LO the
downstream ~;!ope (inclination [), whereupon the hydraulic gradient at exit may be taken to be
(6)

7. It was claimed by Henderson (1956), and accepted by Wilkins, that Laughing Jack Dam would
not have the spillway capacity detennined from Eq. 1. However, the critical depth approach is
still pursued by Stephpnson (1979) who considers observed departures to be due to air
entrainment.
577

Assuming that the velocity in the toe region can be taken to be uniform, and introducing a
factor to allow for the deviation from horizontal,
(7)

wherein V. is computed from Equations (2) and (6). This gives predictions in close agreement
with laboratory results by Sandie using modcls of 19mm aggregate and ~ values up to 90.
Elsewhere within the approach and downstream regions, surface profiles may be computed
by applying the equations of channel flow, which for flow through rockfill take the form
(Parkin et aI., 1966).

(8)

where y = fl( .v depth, ib is channel slope and v is the mean velocity. Eq (8) can, if required,
be integrated with appropriaie substitutions.

4.3 Analysis of Flow Fields

The analysis of seepage flow is normally directed at the production of a flow net, consisting
of an orthogonal set of stream and equipotential lines. In a turbulent flow situation, these
cannot have the same meaning as in laminar flow, but if a streamline is taken
to be a mean or most probable fluid path then equations such as (3) can be applied to Ule
production of a turbulent flow net.

- - Nonlinear FE solution
- . - - Exper free surface
---- Darcy free surface

Fig. 5. Comparison of flow solutions in rockfill dams, with and williout core
(after Volker, 1969).
578

If a finite difference solution is proposed, equation (3) can be combined with the continuity
equation to form the field equation of Lawson, Trollope and Parkin:

where is a scalar potential function with derivatives <\>, (jl, in the canesian directions and N
<\>
= I/o.. This equation can be solved after the manner of Cunis and Lawson (1967).
In general, however, it will be more convenient to usc finite clement methods which can
more efficiently handle irregular boundaries (such as free surface). In this case field equations
can be developed in a variety of ways from equations (3) and (4) or alternatively solutions can
be derived from the methods of variational calculus, which effectively minimise
energy loss. A series of equations for each of these situations is given by Volker
(1969), together with some comparative results as in Fig. 5.
In their development of the variational principle for application to turbulent flow, it was
shown by Fenton (1968) and Parkin (1971) that the total rate of energy dissipation will be
given by

(10)

(where qj = mean seepage velocity, and dv is an clement of volume), and that the flow field
will be detennined by the condition of E extreme, subject to the energy loss equation being
satisfied everywhere:

(11)

From eq. (3), F(V) must take the fonn

F(V) - aN (H'i H'i) (l-N)f2


(12)

whence by substitution in eq. (10) the flow field can be solved from

(13)

where r = (1 + N)/2. The Euler equation, satisfied at the extremum, is the condition of
continuity.
(14)

In solving Eq. (13) by the finite element method, it was noted by both Volker and Fenton
that a streamline boundary is of "natural" fonn, such that conditions are automatically satisfied
in the variational solution, without requiring specification. Typical solutions from Parkin
(1971) are given in Fig. 6 for flow through a homogeneous bank, from which it may be noted
that the elements of a turbulent flow net change geometry with increasing velocity,8 whilst
8. Wilkins (1963) describes a sketching procedure for flow net construction which incorporates
this change of geometry.
579

otherwise remaining orthogonal. Otherwise, the laminar and turbulent flownets do not differ
greatly, except in that the phreatic surface is somewhat higher in turbulent flow.

4.4 Stability

On the presumption that there is adequate security against overtopping, unreinforced


throughflow rockfill structures may fail by either of two mechanisms: the unravelling of surface
stones on the seepage face or internal slip. The former is a progressive development which
will, if unchecked, lead to eventual degradation and breaching of the crest, but the h:.tter is
quick and catastrophic in its consequences, as observed in laboratory models (Parkin, 1963,
Parkin et al. 1966).
From Coulomb theory, it is generally considered that slopes of cohesionless material, such
as rockfill, fail on planar surfaces parallel to the slope (inclination ~), whether wet or dry. If
dry or statically submerged, the factor of safety will be (e.g. Scott, 1980)9

(15)

but in the exit region of a self-spillway dam, it will be

F = Fa iJ(l + iC> (16)

at the top of the seepage face, or if discharging into a tailpond1o


(17)

where in is the hydraulic gradient normal at exit and ic = i/yw' Equation (16) will
generally predict a factor of safety of around 112 F o ' which means that a stable
slope under exiting seepage will be of the order of half the angle of repose for dry
loose rock.

Self spillway rockfill dams are, however, significantly different from an infinite
slope, and laboratory studies have shown that traditional circular slips do indeed
occur in models constructed of uncompacted screenings. These slips generally
follow closely the concave shape of the phreatic surface, and the application of
stability analysis then predicts two critical failure surfaces (the other being for
shallow surface slides). By plotting experimental slips on a classification diagram,
it would appear that the best defensive procedure is to avoid any operating
condition falling within a critical region so defined (Fig. 7). On this diagram, note
that the seepage line represents daylighting of the phreatic line in the crest region,
and that the barrage region is one where the core cut-off is having a negligible
effect on flow. Superimposed on this diagram is a line derived by Johnson (1971)

9. From stability charts ba5ed on a curved Mohr failure envelope, Charles and Soares (1984) found
that F for a compacted rockftll slope decreases modestly with increasing height, and increases by
some 20% with total submergence.

10. Putting cos [3 (normally - D.8) = I, and assuming i,. < about 0.3.
580

Turbulent Flow Net. N = 054

Laminar Flow Net. N e I


with Turbulent Equipatentials as broken lines

Fig. 6. Laminar and turbulent flow nets (Parkin, 1971)

from the progressive overtopping and failure of Hell Hole Dam, California, during
a flood in December 1964. The region above this line could therefore be taken to
give a better bound to the region of danger.
If, in a field situation, a structure is to be formed of compacted quarry-run rock
(as opposed to uncompacted screenings), the relevance Fig. 7 may be doubtful as
there are yet no records of slip failure for such materials in overtopping situations.
However, compacted rock is unlikely to be used in those regions of a self-spillway
dam where good permeability is required.
Slip circle analyses have been performed on overtopped trapezoidal banks of
rockfill, using Bishop's method and incorporating pore-pressures from a turbulent
flow net. These studies by Wilkins (1963) and Fenton (1968), for slopes ranging
from 33 to 45, can be used to define the extent of the zone needing stabilising
reinforcement by drawing an envelope to slip circles having F = 1 (factor of safety).
It has been shown that this envelope can be reasonably drawn as a line parallel
to the downstream slope and in a horizontal distance of 2/3 H (bank height). In
the event that a tail water pond is present, the safety factor on any given circle
increases, but other circles become critical instead, so that there is no net benefit.
However, in the absence ofa stabilising surface mesh, the primary mode offailure
will in all cases be by the erosion of surface stones, in which case slip failures will
not occur.

5. Overflow Rockfills

The design of non-reinforced rockfills for submerged flow conditions can embrace
a range of situations from the construction of cofferdams by dumping rock in
flowing water to rip-rap protection of overflow chutes.
581

i'OO . . . - : - - - - - - - - - - - - - - - - - - - - - - - - - ,
Over topping, H = he + He

Seepage line
075 he
H/He~1+116-
He
I
"uO'50
I .E I~
1-8
025
LLa;
l:
It

Danger of slip
Stone: 9819 mm
---
Hell Hole Dam failure

/<:? f3 ~ 40 Barrage Region

o 05 10 15

Fig. 7. Classification Diagram for Rockfill Dams (Parkin et aI., 1966)

The former of these was studied initially by Isbash (1936), who identified four
stages in the development of the emerging rockfill bank up to stream closure. This
work was later pursued by Olivier (1967), Stephenson (1979), and others, but will
not be taken up here.
Of more importance is the matter of rip-rap design for application to stream bed
stabilisation and overflow chutes, for which the design curves of Olivier (1967)
have been widely used (Fig. 8). From model studies of shallow layers of aggregates
on an impervious base, Olivier identified threshold flow as the point at which local
movements and realignments of stones occurred, but short of total collapse. This
threshold flow (q,) was shown to be substantially affected by the geometeric
arrangement of the surface stones such that the natural packing factor Pc of
randomly placed stones could be varied:t 33% according to whether the stones
were selectively placed flat (unstable) or on edge (stable).
As there were found to be practical difficulties (e.g. determining whether stones
were at or below the surface) in evaluating the areal packing factor, an alternative
procedure was proposed (without any clear support) whereby it could be derived
as the 2/3 power of the volumetric packing factor, which proves to be (1 + e) for
spherical particles (e = void ratio). Threshold flow is then calculated from the
equation l l

11. Olivier considers this to apply equally to the case of an impervious base and to the loe of a
rockfil1 dam where there is emerging seepage flow.
582

(18 )

where K (C/P i/3 = 0.036 for crushed rock (12 to 60mm)


C a constant
ds equivalent stone diameter (m)
Gs specific gravity of stones
s bed slope (tan ~)

Model tests confirmed that ql varies with d/ fl , and equation (18) was then used to prepare
design curves as in Fig. 8 taking Pc = 1.2 (essentially a random tight packing, with no
preferential alignment). Full details of the computation of packing factors are given by Linford
and Saunders (1967).
The application of this procedure to a low rockfill spillway designed to carry some 200
cumecs, on the Broken River, Victoria, is described by Collett (1975). This consists of a 1m
layer of rockfill armouring over a two stage crushed rock filter, and it is of interest to note that
the slope increases from 1: 12 in the upper section to 1:3 at the bottom, where scour is not so
critical because of the build-up of tailwater under high flow conditions.
In designing an armoured rockfill chute, such as the one illustrated in Fig. 9, it is clearly
desirable for the rockfill to be worked over with a backhoe to ensure that the largest stones are
teased up to the surface, and, if possible, also tilted on edge to achieve a low packing factor.
This will normally require some skill.

S 50 OJ
... c:
c:
Cl>
Qj
E
E VI
VI
0
is 20 0
~
OJ
c: OJ
.Q c:
10
I/) E
I/)
C
OJ 05 OJ
"0
>
0
s 0-2 E
0-
a-I x
W 2
0-05 a.
For Pen = 1'2
0{)2 a.
<[
(Packing Factor)

o 10 20
Threshold Flow (eumecs/m)

Fig 8. Design Curves, Flow over Rockfill (based on Olivier, 1967)


583

Fig. 9. Rockfill chute on Blackall Creek, Victoria Australia


(by courtesy LN. Drummond & Associates, Bairnsdale, Vic)

6. Mesh-Protected Rockfills

The first use of mesh-protected rockfill in Australia (apart from the two structures
described by Ash, 1978) was for the Meadowbank cofferdam, Tasmania (1962)12,
with development thereafter taking the course detailed by ICOLD (1986). In each
case, the elements of the reinforcing system consist of a facing mesh to retain
surface stones, which is in turn tied back into the body of the rockfill by a system
of tension anchors.
It has been most usual for the surface element to consist of weldmesh or a grid
of reinforcing bars, in which case the maximum opening for compacted rock is
considered to be about that of the maximum stone size (Wilkins, 1963). In some
cases, a finer mesh has been used, ranging down to 50mm chain wire, but
experience has shown this to be vulnerable to tumbling rocks and logs, which was
evidently the cause of a partial failure in Cethana Dam (HEC, 1969) during
overtopping (and similarly Paloona cofferdam). For this reason, some later
structures included an overlay of heavy bars to protect the primary mesh (PWD
of WA, 1971; Fitzgerald, 1977).

12. An earlier proposal for the use of mesh protection on Wayatinah B Dam (Wilkins, 1956) did not
proceed in the final construction (ICOLD 1986).
584

Fig. 10. Gabion Reinforcement. as used on H.E.C. Pieman River Dams (Cole. 1983)
(Photos by courtesy of the Hydro-Electric Commission of Tasmania)
(a) Upstream cofferdam. Reece Dam
(b) Main rockfill (with anchors). Mackintosh Dam
.'iX.'i

In recent years, the Hydro-Electric Commission of Tasmania has used almost exclusively
stone-fiIled cylindrical gabions, formed of weld mesh and measuring O.9m diameter by 2.4 m
long. These are considerably less sensitive to surface damage. For cofferdams, where tension
anchors are not normally provided, the gabions are stacked with their axes aligned in the stream
direction, which provides greater stability. However, for main embankments such as Reece (or
Lower Pieman) Dam (Cole, 1983) the gabions are aligned across the face cf the dam, for
economy and better appearance but incorporating, in this case, a system of internal anchors
(Fig. 10).

-~----~

Ww

,
Hw-,J---K1

-l--
~~QQ~~~~~

I __ L
~ u R2

Fig. II. Forces on a Sliding Wedge (Wilkins, 1963 Guidici, 1967)

The design of an anchor system is normally based on a sliding wedge stability analysis, such
as that of Wilkins (1963), which assumes the force system shown in Fig. II a. Because <' for
many compacted rockfills is close to 45, the reactions R j and R2 are assumed to be co-linear,
so that

F ~ (W - U) tancl>' (19)
H w -h w

where F is the factor of safety of unreinforced rock fill. It is then nccessary to provide
sufficient steel (reducing the denominator) to bring F to an acceptable value (eg, F= 1 on
normal working stress). The same calculation wiIl also determine the extent of the region
requiring reinforcement, from the analysis of a series of trial wedges.
Minor modifications to this procedure were introduced by Guidici (1967), based on the model
in Fig. 11 b, which provides for di fferent depths of overtopping and different water profiles
down thc slopc. It was found from the model tests of Wilkins that the pore water forces Ul ,
and U2 fitted an equation of the form

u ~ a L" y w (20)
586

with n '" 1.6, and curves were produced for different degrees of overtopping, specified by t =
o and 3m. Active pressure was calculated from Rankine theory, and for the evaluation of F,
pore pressures were factored up by 1.2 and tan $' factored down by 0.9. The volume of
rockfill to be reinforced is then given by the length L to give F > 1, which is generally in the
region of 2/3H in conformity with previous slip circle analysis. Other refinements to this
procedure are possible, such as to determine p. from a Coulomb wedge analysis, as described
by ICOLD (1986).

r
1.3
~1 Surface grid of
20 mm bars
Anchors at 1.2 m hor. SpaCing; 1.2 x 0.3 m

IE
I CJ)
I .
'0
~i_~ _ _ _ ~_

BORUMBA DAM

1.3

Hor. bars at 1.35 m


Anchors at 1.5 m hor. spacing - \ . .... ______ vert.spacing
. . . . . . . . . .. "" ' /~
'J' / //'
' .....~. / '.</~/;~//>?,
.' / T,..' /
/ r- 50 mm Chain

'- c='- c-_-,_- _S~T-A~G_-_E~2~_~ _.~/ v~>~;:~;,6<~:;"


'_' ' mosh

ROWALLAN DAM

Anchors at 1.8 m vert. 50 mm Chain wire mesh


x 0.6 m hor. / - Overlay of 20 mm bars
, at 0.3 x 0.9 m

1.75
~1
TOONUMBAR DAM

1 no. 24 mm Unformed - - - - , 0.94 x 2.4 m


bar/gabion - concrete cylindrical gabions

I ~~
I
I~I--I----------~~~~:~~~~~~~_~_~~-~~~~_J-r~
1.3

I.... '-24 mm dowel \\ \\.


~1

grouted into rockfill MURCHISON DAM

Fig. 12. Tic bar arrangements for reinforced rockfills (As referenced in ICOLD (1986),
except Toonumbar Dam for which see WC & IC of NSW (1970
587

Whilst cofferdams will nonnally be provided with full face reinforcement, on a main
embankment it will be possible for the reinforcement to be tenninated at some appropriate
level. This is known as the mesh protection level, which must be established from hydrological
and constructional considerations to give the desired degree of protection (ICOLD, 1986).

6.1 Bar Spacing and Configuration

Vertical bar spacing will nonnally be selected so as to ensure that surface bulging of the mesh
is kept to a minimum, but may also be influenced by other factors such as rock size and
construction methods (notably Meadowbank). Wilkins (1963) considers this spacing should
not exceed half the bar length, with a further 1.5m allowance on length (with hook) for bond
development. From the examples reported by Fitzpatrick (1977), it would appear that bar
spacings are generally I to l.5m, and that the only instance of unsatisfactory perfonnance from
a too large bar spacing (4.5m) was thc Meadowbank cofferdam. Bar configuration has also
received some attention, with the objective of ensuring that the uppennost level of crest
reinforcement will be adequately anchored back into rockfill at all times, to guard against
progressive degradation of the crest in a sudden overtopping. This has led to the sequence of
developments detailed in Fig. 12, involving the use of cranked and inclined anchors, and the
use of grouted dowel anchors.

6.2 Perfornlance

The failure of Cethana Dam (referred above) occurred when the main bank had reached a
height of 15 m and was required to pass a flow of 170 cumecs. Because the uppennost layer
of rock could not be protected in advance, (which is one of the advantages of gabions), this
was progressively eroded over some five hours, until a sudden breach opened on the left side,
apparently due to perforation and loss of the surface mesh. The resulting scour was 9m deep,
representing the loss of 15000 m3 of rockfill, but elsewhere the dam showed no evidence of
any distress. In the case of Leichhardt Dam (NW Queensland), overtopping in an 850 curnec
flood caused some failure and loss of rock (Davey, 1960), but elsewhere in Australia
overtopping flows have been carried with complete success (eg. Robson, 1965; Fitzgerald,
1977), as in the illustration of Googong Dam carrying a 2.5m overtopping flow. These include
some quite substantial flows, as at Ord River Dam (northern Western Australia), overtopped
continuously for six months at depths of over 10 m and estimated flows up to 5700 cumecs.
In anticipation of surface velocities up to 8.5 m/sec, the part-built embankment was annoured
over thc top and downstrcam faces with a 1.8 m laycr of 0.9 m rocks, covcred with light mesh
and an overlay of heavy bars (PWD of WA, 1971), but suffered no distress beyond local
settlements of rockfill under the mesh of up to 0.3 m.

6.3 Protection of Cohesive or Impervious Fills

In cases where a relatively impervious fill or natural fonnation is to be protected against


overtopping, a mat of rectangular gabions has been found to function satisfactorily, with little
or no assistance from internal tics. Following the overtopping of a spillway chute at Baroota
Dam (ncar Port Pirie, South Australia) in 1974, the underlying highly fractured and deeply
weathered siltstone eroded to such an extent that the integrity of the dam appeared to be
588

Fig. 13. Googong Dam overtopping in 20 year flood, 16 October 1976


(Photo courtesy Canberra Times)

threatened (Johnson, 1977). For the purpose of a temporary repair, fill was placed with hand
wackers to re-shape the eroded gully, and covered with filter cloth and a mat of 2 x 1 x 114m
rock-filled gabions. These were wired together and onto 20mm horizontal bars, and 20mm
stakes were driven through the mat into the fill and foundation to provide some measure of tie-
down. The lower end of the mat was left free to deflect in response to erosion downstream.
The arrangement was tested soon after in a 10 year storm, which produced a flow of 125
cumecs (approx. 6 cumecs/m width) and a velocity up to 10 m/sec. Despite appreciable
shifting and compaction of stone in the gabions, exposing the filter cloth, no damage to the
underlying formation occurred.

6.4 Permanent Hood Protection

As all the abovementioned cases required only temporary flood protection, corrosion of the
mild steel armouring was not an issue. In some cases, However, it may be required to function
over a longer term. In the case of Moochalabra Dam, supplying water to the town of
Wyndham, Western Australia, stage construction was proposed, with the stage I structure being
required to pass floods by overtopping (Wark et a!., 1982) (Fig. 14). It was therefore necessary
for the entire mesh and anchor system to be galvanised, in which state it has functioned
satisfactorily since completion in 1972, being overtopped for 3 to 5 months each year by water
depths up to 1.25 m (around 300 cumecs or 2.5 cumecs/m). This structure can thus be
regarded as a development of the original self-spillway dam (Laughing Jack), in which the
flood capacity has been increased around ten-fold. Possible alternatives to galvanising might
inelude a corrosion allowance on mild steel (3 to 6mm) or the use of aluminium, factoring up
the cost by about 2 and 5 respectively (Wilkins, 1963).
589

Horizontal

HEIGHT
15m

12

Steelx 150
~r--:~~~~~~,,-'50 fabric under
x 76mm
9
3A surface bars

6 38 All bars 25mm 0


Notural
surface LONGITUDINAL
~--~----~"''' -----
BARS IN HOOKS

IJ
o
! ,
5
!
10
TOE ANCHOR
! !
AT 1-8m ClC
METRES

Fig. 14. Moocha1abra Dam - Typical section (Wark et al. 1982)

Apart from Moochalabra, the only use of permanently reinforced rockfill has been on two
estuarine barrages in Northern New South Wales, Australia, in service since 1960 (Ash, 1978),
and on the Pit 7 Afterbay Dam, California, built in 1965 to regulate power station discharges
into the Pit River (Schackleford et al., 1970). The latter is 11 m high, and carries normal flows
up to 180 cumecs and extreme flows up to 2400 cumecs.

7. Conclusion

From unfortunate beginnings, the technology of passing floodwater through and over rockfill
has advanced to the point where such flows can now be accepted confidently on either a
temporary or permanent basis.
The se1f-spilway dam, as conceived by Wilkins and culminating in the construction of
Laughing Jack Dam, was originally based on unrein forced rock fill and restricted to rather low
flows, of the order of 0.5 cumec/m. Subsequent experience on Moochalabra Dam has shown
that such capacity can be substantially increased (to about 5 times) by the use of a reinforcing
mesh, and that this can be made to function reliably over a long term. If the structure is to be
unreinforced, then it must be established that the phreatic surface under the design flow can
be safely accommodated. For this purpose, it has been shown that the head-discharge law is
linear, that the phreatic surface in turbulent flow is higher than for Darcy flow (for similar head
differential) and that the turbulent flow net has clements of varying dimensions (LIB ratios).
The more widespread application is to temporary protection of cofferdams and rockfills
during construction, based on the use of mesh and anchor systems. In present times, this
almost invariably involves compacted rockfills, wherein slip failures are, at this point,
unknown: failure, if it occurs is by unravelling or erosion. Of the various protection systems
reported, the use of anchored chain mesh as at Cethana Dam, could be regarded as not
satisfactory in that it cannot be secured quickly enough in an emergency and is too much at
risk to perforation. In contrast, wire gabions have been shown to produce a very robust
590

structurc, albeit at a price. With appropriate protection, however, it has been shown from many
case histories that rockfills can be rendered stable under even quite massive flows at a cost far
below what would otherwise be required in alternative diversion arrangements.

Acknowledgrrnrnt

Compilation of this report has been much assisted by reference to material assembled by my
colleague, Professor Jack Lawson (Lawson, 1987), and by contacts over many years with Mr.
John Wilkins (formerly H.E.C.). These contributions are gratefully acknowledged.

References

Anon. (1989). The Briseis (Cascade) Dam Failure. Australian Geomechanics No. 17 (June),
p.19.
Ash, R.R. (1978). Discussion on M.D. Fitzpatrick, ANCOLD Bulletin 50 (April) p.23,
(Australian National Committee on Large Dams).
Bakhmeteff, B.A. and Feodorov, N.V. (1937). Flow through granular media. Trans. ASME
(Series E), Journal of Appl. Mechs., V.A. pp. 97-104.
Barnett, R.H.W. (1990). Discussion of Cascade Dam failure. Australian Geomechanics, No.
18 (January), p.ll.
Charles, J.A. and Soares, M.M. (1984). Stability of compacted rockfill slopes. Geotechnique
34:1, pp. 61-70.
Cole, B.A. (1983). Concrete faced rockfill dams in the Pieman River Power Development.
Inst. Engrs. Australia, Civil Engineering Transactions, CE 25:3, pp. 162-169. See also:
Knoop, B.P. and Lack, L.J. (1985). 15th Congr. Int. Comm. on Large Dams, Zurich, 1:1103-
1120.
Collett, K.O. (1975). Unusual surfaces for large spillways. ANCOLD Bulletin 42 (July), pp
3-10.
Curtis, R.P. and Lawson, J.D. (1967). Flow over and through rockfill banks. Proceedings,
ASCE, v. 93 No. HY5, pp. 1-21.
Davey, GJ. (1960). Rock fill dams at Mary Kathleen and Mount Isa. Journal, Inst. Engrs.
Australia 32: 12 (Dec.) pp. 291-300.
Fenton, J.D. (1968). Hydraulic and stability analyses of rockfill dams. Report DR 15, Dept.
of Civil Eng. University of Melbourne, Australia (July).
Fitzgerald, B.J. (1977). Rood forecasting during construction of the Googong Dam.
ANCOLD Bulletin 47 (April), pp. 13-15. See also: ANCOLD Bulletin 49 (October), pp.27-
40.
Fitzpatrick, M.D. (1977). Reinforced rockfilJ in Hydro-Electric Commission dams. ANCOLD
Bulletin 49 (October) pp. 20-26.
Giudici, S. (1967). Discussion of Parkin et al. (1966). Proceedings, ASCE, vol. 93 no. SM5,
pp. 329-336.
Henderson, F.M. (1956). Discussion of Wilkins (1956), Proceedings, 2nd. Australia - New
Zealand Conference on Soil Mechanics and Foundation Engineering, Canterbury, N.Z., P.
148.
591

Hydro-Electric Commission, Tasmania (1969). Cethana Dam. Flood breach of partly


completed rockfill dam. ANCOLD Bulletin 28 (July), pp. 22-36. See also: Mitchell, W.R.
(1973) River diversion arrangements for the Cethana Power Scheme. 11th Congr. Int.
Comm. on Large Dams, Madrid, vol. 2, pp. 145-157.
ICOLD (1986). River control during dam construction (Bulletin 48a). Appendix 1 by
Australian National Committee: River control during construction, pp. 131-164.
Isbash, S.V. (1936). Construction of dams by dumping stone in running water. 2nd Int.
Congress on Large Dams, Washington D.C., Vol. 5, pp. 123-
Johnson, H.A. (1971). Flow through rockfill dam. Proceedings, ASCE, 97:SM2, pp. 329-340.
Johnson, N.A. (1977). Temporary protection of Baroota spillway using gabions. ANCOLD
Bulletin 48, pp 24-26.
Lawson, J.D. (1987). Protection of rockfill dams and cofferdams against overflow and
throughflow - The Australian experience. Inst. Engrs. Australia, Civil Engineering
Transactions, CE29:2, pp. 138-147
Lawson, J.D. Trollope, D.H. and Parkin, A.K (1962). Some hydraulic aspects of
unconventional rockfill dams. Proceedings, 1st Australasian Conf. on Hydraulics and Fluid
Mechanics, Perth, W. Aust., pp. 159-172, (pergamon - Ed. R. Silverster).
Leps, T.M. (1973). Flow through rockfill. In: Embankment-Dam Engineering, Casagrande
Volume (R.C. Hirschfield and S.J. Poulos, Eds.), pp. 87-107.
Linford, A., and Saunders, D.H. (1967). A hydraulic investigation of through and overflow
rockfill dams. British Hydromechanics Research Association, Report No. RR 888 (58pp).
McKeown, M.R, (1938). Sluicing for tin at Briseis, I and II. Chem. Eng. and Mining Review,
Vol. 38 pp. 385-391 (11 July) and pp. 438-445 (10 August).
Mattes, G.H. (1951). Discussion on Weiss, Transactions, ASCE, V.116 pp. 1175-6. (See also:
Wilson and Squier (1969). SOA Report, Earth and Rockfil1 dams, 7th Int. Conf. Soil Mech.
& Found. Eng. 3:137-223).
Olivier, H. (1967). Through and overflow rockfill dams - New design techniques.
Proceedings, Inst. of Civil Engineers, London, Vol. 36 pp. 433-471 (paper No. 7012).
Parkin, A.K. (1963). The hydraulic and stability characteristics of self-spillway rockfill dams.
M.Eng.Sc. Thesis, University of Melbourne, Australia.
Parkin, A.K (1971). Field solutions for turbulent seepage flow. Proceedings, ASCE, v. 97
No. SMI, pp. 209-218.
Parkin, A.K., Trollope, D.H. and Lawson, J.D. (1966). Rocklill structures subject to water
flow. Proc. ASCE, Vol. 92, no. SM6, pp. 135-151.
Public Works Dept. of WA (1971). Overtopping of stage I of the Ord River Dam. ANCOLD
Bulletin 34 (September) pp. 3-6. See also: Webster, KC. (1973). Spillway design and river
diversion for the Ord River Dam. 11th Congr. Int. Comm. on Large Dams, Madrid, Vol. 2,
pp. 553-564.
Robson, W.M. (1965). The passage of floods during the construction of Borumba Dam.
AN COLD Bulletin 17 (October), pp. 14-16. See also: Learrnonth, F.M. and Butler, N.J.
(1967). Surface diversion of streamflows .... , 9th Congr. Int. Comm. on Large Dams,
Istanbul, Vol. 2, pp. 853-876.
Sandie, R.B. (1961). A laboratory investigation of self-spillway rockfill dams. M.Eng.Sc.
Thesis, University of Melbourne, Australia.
Sarkaria, G.S. and Dworsky, B.H. (1968). Model studies ofarrnoured rockfill overflow dams.
Water Power 20 (November), pp 452-462.
592

Scott, c.R. (1980). An introduction to soil mechanics and foundations (3rd Ed.), Appl. Sc.
Publ. pp. 406.
Shackleford, B.W., Leps, T.M. and Schumann, J.E. (1970). The design, construction and
performance of Pit 7 Afterbay Dam. Transactions, 10th Int. Congress on Large Dams,
Montreal, Vol. 1, pp. 389-404 (see also: Leps, 1973).
Stephenson, D.A. (1979). Roekfill in hydraulic engineering. Developments in Geotechnical
Engineering, No. 27 (Elsevier), 215 pp.
Volker, R.E. (t 969). Nonlinear flow in porous media by finite elements. Proceedings, ASCE,
V. 95 No. HY6 pp. 2093-2113.
Wark, R.J. and Szymakowski, 1. (1982). Moochalabra Dam - 10 years' experience with an
overtopped rockfill. ANCOLD Bulletin 63 December), pp. 40-43.
Water Conservation and Irrigation Commission of NSW (1970). Two dams in northern New
South Wales. ANCOLD Bulletin 30, pp 52-54.
Weiss, A. (1951). Construction technique of passing floods over earth dams. Transactions
ASCE V.116 pp. 1158-1173, Paper No. 2461. See also: Author's closure, pp. 1177-8.
Wilkins, J.K. (1956). Flow of water through rockfill and its application to the design of dams.
Proceedings, 2nd Australia - New Zealand Conf. on Soil Mechanics and Foundation
Engineering, Canterbury, N.Z. pp. 141-149.
Wilkins, 1.K. (1963). The stability of overtopped rockfill dams. Proceedings, 4th Australia -
New Zealand Conf. on Soil Mech. and Found. Eng., Adelaide, Australia, pp. 1-7.
Wilkins, J.K. (1989). Rockfill - a brief history. Australian Geomechanics No. 17 (June), pp.
7-8.
CH APTER 19
SPECIFICAT IONS AND CONTROL OF NATURAL ROCKFILLS
H. EVRARD

1 - I NTRODUCTION

For both tec hnic al and e conomic reasons. "natural" rockfi ll s obtained
from large r ock quarries have bee n u sed for centuries to fi ght coast-
line erosion , to prote ct river ba nk s and in building earth dams.

1 - 1 - The technical context

There a r e real benefits from using this type of material : build - up is


facilitated and relatively t r o ubl e-f ree. Per linear meter , dykes use
mo r e conc rete tubes thsn natural materials due to the shapes of the
rocks and the ways in whic h they can be 8ranged. Other prob lems related
to concrete cubes have been re por t ed including difficulties in obtai-
ning high densities and alkaline r eactions.

1 - 2 - The economic context :

Natural rockfills are often supplied from g r avel or dressed stone quar-
ries wher e the rocks conce rned are con sidered unworkable as they cannot
be c ru shed or are too irregular in shape. In these conditions , ex-
quarry prices can be relatively low. The preparation of concre t e blocks
implies sourcing gravel and cement whi ch in some countries are not al-
ways available. The higher density o[ "natural rockfill obta i ned from
heavy rocks vs. con crete a l so means that significantly less tonna ge can
be used to meet no r mal scal ing standards in whi c h hardness is a major
fractor. Economic studies s hould not therefore only consider pr i ces per
ton.

However, natural r oc k fills are not defect-free. Rocks may be cracked,


and thus sensi tive t o wa ter and frost. Fo r thi s rea son it is vital to
define the st resses and st rain s to which the ma ss will be subjected wi -
thin the st r uc t u r e and to apply specifications to ensure that the rock-
fill meets i t s objectives. Another consideration is that major
cons truction projects rnay require grea t er quantities of roc kfill than
are available from local quarrie s . This may imply prospect ing for new
local sources, estimating preCisely the quantities they may contain and

'"
E. Moronho dos N,,u (<,d.J. Adl'tJIIUS III R{)("lifill S'n/("II/r<,s. 593-609.
C 199 1 KJ" ...."r An,demk Publishers.
594

finding applications for by products which in some cases represent 80%


of the rock extracted. If the usable percentage is not evaluated
correctly, the result may be catastrophic, particularly if the problem
is only discovered when construction work has already begun.

All these topics are discussed in detail in a report on rockfills pu-


blished by the Ponts & Chauss~es Central Laboratory (see Bibliography
below). This paper is limited to considering essential specifications
and controls to apply to supplies and applications.

2 - SCALING ROCKFILL REQUIREMENTS

Construction projects are designed to be economically and technically


viable. For this reason, several variations to geometric and hydraulic
plans are prepared for every project. Scaling is decided by hydraulic
engineers who use all aspects of environmenal data to determine the de-
sign parameters. For example, seawall structures are designed to meet
specific maritime problems such as tides and currents and river embank-
ments to solve problems of flow rates or to protect construction sites.
Sea wall and embankment composition is generally a compromise between
two often contradictory preoccupations

- to use sensibly priced materials

to ensure adequate protection of construction work throughout the


project with an acceptable lifespan.

Usually, dyke structures include (figure 1)

- a basement wall (between the natural ground and the foundations of


the construction

- a main structure made of up several layers (filter, core, etc ... )

- a shield wall which bears the brunt of the assaults and can be made
up of one or several layers of rockfill.

The scaling of every part of the structure starts with calculating the
mass (or diameter) of the unit blocks of the materials to be used in
the construction. The most commonly used formulae are empirical and ge-
nerally based on experimental studies carried out on small-scale mo-
dels. All formulae include hydraulic requirements.

Figures 2 & 3 below give some of the formulae, in general use for cal-
culating mass
595

r igure 1 Examples of applications in rivers

I/o 1,5 Phen +5


---+-=w---+ ~
Grof'lie roci<'''1

Roci<fill o:;~
top soil Rhine rtver embankment

Quarry run Intill


o

Fabnc filter
TO -soil 020

Random bad+- fill ~.~~.Ji/~wal!!r mori<


. - . - . - Coo/mat limit
IClise fiVer embankmen l I
Ror:kflll +20"/0 0/60 "'-d~~~=-- Rock fill
'TO ve gruvel >.:""".~~~ __ ~...2! the rIVer Oise

Random
IDam built In severd. s Ktions I

IExamples of natural rocJdili applicatIOns I

----~
- -..=.--
596

Fig. 2
Common formulae and notations used by
hydraulic engineers (maritime applications)

LARRAS'S Formula SOCREAH Formula

Block mass Block mass

w .,-__C"rd_H_'_P.:.._ __ (1 ) w. p, kH'( I -OIS') ( 2)


(~ - i )'.'0> 4,D (~ _ I)' (0(1 a - 0.8 .
Q - )Or. ai' un>.. - - - '
L

(for the shield wall) where k 0.25 for quarry rockfill


k 0.12 for concrete cubes

(for the shield wall)

R. HUDSON'S Fon=ula R. HUDSON'S Simplified formula

R. HUDSON'S Formula and tables (see below) P H' H


are widely used (4 )
p - 1)' COll:t
(~
h
Block mass

where h - calculated rockfill dimensions


expressed in terms of the intermediate
(3)
COll'l. level in meters
(for abutment)
N.B. - For the SOGREAH, the largest di -
(for the shield wall) mens ion of the abutment must be a maximum
1.2 or 1.3 times the trough of a breaker
below the PBBH level.

Definition of the notations used in scaling formulae by hydraulic engineers.

w~ 114" It I
II : .vu.~. IU.I
Ie
H I taaku !'It:gl'1t tal
... Hil Buutr up~ 1 :ud.
~r dtpth of the bioCIl HI. tera, of depth of the .... at.r (_I

a t;.&r.. to hNllor.ul &f\Q1I ,oJ


)J.CC.ttlCUr:.t of frlctlon
vOld t.tlO
CI. t;lloci: Ih.ap.e coe!!H:1I..ot

P.tlock dt!",;llt.y It/a3)


fI",wl":.tr da.r.llty tt/&3)
Kaltabl11ty tltlO ~Ied In the fOr.\;l ..

Shald 11 th.1Ck..n ...

If. "ua!l4r of bloCkl per .urfice una

Q ~l.or 11"" Uti lall

k. rou;t-.r.... factor. w.t flow nct lOlL

d dJ&.Nter of I IPlitrlCll rod


597

Figure 3 : SOGREAH formula defining empirically the mass of a rockfill


in a river envionment

Current velocity (m! s) Mass (kg)

3 29
3.5 74
4 164
4.5 331
5 625
5.5 1107
6 1870

1J - V6
25

where V is current velocity near the rockfill


W is the mass

It is impotant to emphasize the difficulties encountered to homogenize


the various formulae, some of which are empirical, derived from more or
less recent bibliographies. There is currently much terminological
confusion between weignt and mass. To remain in conformity with the In-
ternational Systeme (IS), we will consistently use the term Mass in the
formulae. It should be noted that the main rockfill scaling formulae
include the notion of a material's theoretical or true density in thr
form of the following expression :

-r
Pa

(~:
where Pa is the true density of the material in t/m~
Pw is the density of the water (1.026 t/m 3 in maritime
environments)

The Table in Figure 4 below shows the fundamental role of the density
when scaling rockfill blocks

Figure 4 - Function of Pa in scaling formulae

Pa Pa
( t/m3)

(~: -1) 3
3 0.42
2.7 0.S2
2.6 0.72
2.5 0.84
598

To obtain an eqivalent shield wall stability it is necessary to in-


crease block mass when Pa decreases for example by

100% when Fa falls from 3.0 to 2.5 t/m 3


35% when Fa falls from 2.7 to 2.5 t/m 3
16% when Fa falls from 2.6 to 2.5 t/m 3

This quite naturally leads engineers to design projects with the hi-
ghest theoretical Pa possible, almost always above 2.6. This can lead
to absurd situations if local or regional rockfill resources are incom-
patible with this hypothesis. Design offices calculations must inte-
grate the properties of locally available rockfill. Coefficient kd is
also a highly significant factor as it varies, for example from 2.1 for
rounded rocks to 3.5 for pointed rocks and 7.5 for concrete cubes.

All these factors have to be considered in the specifications for sup-


plies and supply ~ontrols.

3 - RECOMMENDATIONS FOR SPECIFICATIONS

As the rapid discussion of rockfill scaling in the previous chapter


shows, hydraulic engineers consider rockfill mass and shape to be vi-
tal. Both parameters are used in scaling formulae and Contractors re-
quire blocks to be in conformity with these specifications ; once posi-
tioned they must remain intact whatever the stresses and strains ap-
plied.

It is therefore essential to ensure that

1) rockfill is crackfree

Cracking is usually directly related to the structure of the


vein from which the rock has been quarried

2) rockfill can resist the stresses and strains to which it will


be subjected, i.e. :

- abrasion (rocks must not become rounded)


- shocks
- frost in some regions.

All these properties are easy to verify in the quarries concerned.


Contractors must make sure, far upstream from the construction site,
that the quarries selected can supply quality material in the dimen-
sions required and to the construction site schedule (quarry approval
registration). It is also essential to possess adequate equipment to
meet the construction site's requirements (i.e. for sorting, transpor-
tation, storage and use).
599

3 - 1 - ROCKFILL DENSITY

Rockfill density partially depends on rock quality (compactness,


uniformity, cracking) but rather more on its mineralogical behaviour. A
few examples are given below (Figure 5).

Figure 5 : Mineralogical composition and true density of the main types


of rock used in rockfills

Ma in Rock Examp les of rocks Dominant Mineralogical Pa of the rock


(ategar ies Composition (r/m3)

Siliceous rocks Quartzites Quartz 2.55 to 2.65


Gres

Carbonaceous Chalk (used in rockfills) Ca lc He 1.00 to 1.20


Rock 50 Limestones Ca lc He 2.30 to 1.70
Dolomites llmestones lea lc ite Oolomlte 1.70 to 1.90
Dolomites Dolornlte 1.70 to 2.90

Plutonic rocks Gran Hes Quartz


Micro-GranHes Feldspars 2.55 to 2.75
Micas
i
I
Dior He I Feldspars ... amphiboles I
I
I
Micro~Diorites 2.75 to 3.05
Gabbro, dolerites, ophHes Feldspars + pyroxenites

Va leanie rocks Rhyo 1 He Granlte Quiv. 1.60 to 2.70

i Andesite Diolrite equiv. 2.60 to 2.78


Ba sa It Gabbro equiv. 1.70 to 3 10
I Dac ite 1.60

Metamorph ic
I Schists, micas I Micas quartz + feldspars 1.65 to 2.85
m1caceous rocks Schists, hornfels, gne i ss
-~

Amph ibo 1 ic Amph ibo 1ites 2.80 to 3.10


I ",,",""", , "' '"'""'
Pyrox.en ite s Pyrox.enes t feldspars

Damaged or cracked rocks can have high densities (i.e. over 2.6 t/m 3 )
and thus density is not a sufficient selection criterion for rockfill.
Density does not characterize absolute quality but scaling and Contrac-
tors should modulate density to suit local resources and block position
in the structure. We suggest specifications based on quality proper-
ties.
600

3 - 2 - ROCKFILL PROPERTIES

It is important to distinguish between the intrinsic properties of ma-


terials and properties stemming from production processes and handling.

Intrinsic Properties are measured in lodes or quarries.

Supply Properties are inspected as close to the structure as possible.

A further set of properties. called "additional properties" can be


equally fundamental in terms of the environment and specific conditions
relating to the structure (frost-cracking, for example).

Rockfills should thus be divided into 3 classes of properties

- Class Ai and A'i


To quality rockfill

- Class Bi and B'i


Medium quality rockfill for use in averagely stressed zones

- Class Ci and C'i


Rockfill for use in unstressed zones

3 - 2 - 1 - Specifications for intrinsic properties

Cracking is one of the vital aspects of rockfill studies and we have


developed our analysis from an approach to "rock continuity" suggested
by DENIS A., PANET M TOURENQ C. (1979) wich measures the rate at
which a longitudinal wave is propagated and the expression of the
continuity index :

Ic = 100 VIm (observed)


VIc (theoretical

1m

Porous environmMts
/

environments

20

Porosily

Ag.6_ Variation of continuity inct:.x Ie as a function of porosity n for porous


and crocked rocks.
601

I/) 50
Porosity n ~o
Fro51 cracking flsks

IFig 7_ Nonog~am"determine the conlmudy Ind~x, and


degree d cracking de rod< based on IdentlflroilOn
or the roeX
Vi
Ic =-'-100
VI
~re VI =ra'e a' which longl tudlool waves propaJate In the rod< '---_ _ _ _ _ _ _ _---,
Vi=rute atwic;h lonql~qJf)al W?V'i'? propagate calculate-d, from" Ihe I
mrnero(oqlcG{ com12os/,fOn 0 the- rock and m eon raf~ '_'1 mtnera?

For class Ai, Ic is > 80 and the degree of cracking < 20

For class Bi, Ic is > 70 and the degree of cracking < 20

For class Ci, Ic is > 60 and the degree of cracking < 20

In particular, it should be noted that porous rocks are acceptable in


all the above classes provided structures are not on sites liable to
freeze-cracking. The only effect will be on Pa.

We further suggest the use of the Deval test to evaluate thresholds


when using soft or damaged rocks as these tend to become blunt or roun-
ded under the action of breakers loaded with sand or pebbles, for
example. This test (French norm NF P 18577) provides values for the re-
sistance to abrasion due to reciprocal friction under water between the
elements of any specific material.

To summarize the foregoing section on intrinsic properties (as measured


in laboratory)

- density Contractors are free to choose from locally available


resources

Laboratory controls confirm that the quarries selected meet specifica-


tions
602

Class/Trial Continuity Cracking Deval Micro-Deval


index Degree wet with water
1t Of D8 or MOW

Ai > 80 < 20 ~
- 5 16)

Bi > 70 < 20 ~ -4 20}

Ci > 60 < 20 ~ - 3 27)

Additional properties

Porosity : good freeze-cracking risk indicator

n < 2% will not freeze-crack

n 2 to 5% rock unlikely to freeze-crack

n > 5% freeze-crack risk must be controlled

3 - 2 - 2 - Specifications for properties of supplies

For a facies defined using the approach outlined above, the specifica-
tions are as shown in the table below (Fig. 9).

Fig. 9 - Properties of supplies

Class Granularity Rocl<fill shape Continuity index


(defined by dimensional ratio measured on rockfil1
Contractor) .::....!...Q Ie 3 directions on
* 2E site

Af Category 1 < -2 > - 75


Bf Category 2 < - 2,5 > - 65
Cf Cat.egory 3 < - 3 > - 55
* N.B. : the DPNVN prefers to related ration

1 to ~
E 2E
Example : it would be recommended by a rockfill shape

~ approx. 3 with 1 < 3


2E E
603

- Granularities are defined by Contractors to meet stability and filte-


ring conditions

- Shape is an important factor in scaling formulae (i.e. coefficient


kd). It is defined as follows.

Fig. 10 - Dimensional ratio

L the biggest dimension

G the biggest dimension which can be measured perpendicular to L

E the biggest dimension which can be measured perpendicular to plane


P (L.G)

The dimensional ratio is usually determined by ~


2E

4 - CONTROL OF THE INTRINSIC PROPERTIES OF THE ROCK

Intrinsic rock properties are controlled regularly in the quarries as


cutting faces progress (see material and lode date sheet).
604

o ;::;GRGA T~S

o
[] i<:OCKf ILL

;"~E~AGE ;"NtH.:AL QUARRY PROOliC710N

:~C~~Sl'Jf cr ALL TYPES OF SlONE

ROCK, ILL PR::xJUCTlON


Averd-:';E annual fodf111 productl0rt : ,. 5 t l..J....U.1...L tons/year
"2 t Ll.L.L.LJ.....J tar,s/year
> It ~tonsi)ear

;> 0,5 t ~tons/yeM

Usual tlloek size:

Fig. 11
Usua 1 b lock shape: cubic r=l
tabu lar c::J
roundee c:::J Quarrying data sheet

L.:....
1E

SUPPl'i CHARACTERISTIC
- vusa 1 appra lsa 1 of loIhO Ie 0
- rockf j 11 stored 1n 10ne d~rMaed 0
nan), ddlMQed 0
yes

crak ina
0 0
- continuity lnde)l On ~hole rockfl11 C*)
- general rodfn1 class
(*) tests attaChed
"0 BfO orO
ROCKFILl ST()(AGE AREA
- Storage area (opacity L.LJ...J arres
- Quant He In stock. on LU 19 LLJ LJ....L..1....l tons
- ~as a pre-sort imple.ented 1 0,,, 0 eo
- Sorting methOd used?

STORAGE AND LOADHfG EQUIPMENT


1) (It the Quarry-f.~ce

I 1
2) for truO loading 10,'" DO
bucket scoop
1grab DO
scoop
DO
dragl1ne
DO
EQUIPMENT

o o
~EIGHIHG yes '0
- scoop with weigher

o o
'ilr~b

- welghbrldge

QUARRY BLASTntG (optlonel) Prep~rat ton date :


Mdln points of rockfl11 blasting pl~n : Slgrlaturt< !

- surface c;~ered: 112


rdt 10 Spread:
- Bed: Job funct Ion:

- exp los he usee;


- spec If Ic charge:
LP( forI!
605

Fig. 12 - Material and intrinsic properties data sheet

I - ADMINISTRATIVE DATA:

QUARRY NAME
ADRESS
OPERATOR NAME

READ OFFIcr ADDRESS

TEl [PHONE/TEL EX

QUARRYNG PERMIT ESSUl DATE DURATION


f---------------------- ------.---
1 - UETAIL\ ABOUT THE LOGE

Brief sunnary of quarry geology and facies present on


quarry facE!

Tectonics (faults, cracks)

Orientctl0n of layers, thickness of seans (sedlmentary


and metamJrphic rOCks)

(t'.anges In rock cross-section

3 - GEOTECHNICAl CHRACTERISTICS OF WHOLE TYPICAL ROCKS

IDENTIFICATION

INTRINSIC CHARACTERISTICS

ADDITIONAL CHARACTERISTICS

PETROGRAPHY
Dens It} ....'ave propagat ion
- Deval
ReSistance to
Speed V 3 axes m/s
Wet poros; ty frost

MINERAL
- t/m3
606

Fig. 13 - Operation and Supply data sheet

SUPPl Y DATA SHEET

SITE Inspect 10n site:

Operator

Type of supply:

1 Intrlns1c characteristics
Vlsua 1 aspect of materia 1 In conformity with spec 1f teat Ions yes no
1f NO, " of material presumed to be wrong:
Charactertstics of these materials : Pe : o o
lew:
Oh :
n% :
frostcraklng :
Cd leu late theoret lca 1 propagat 10n rate :

2 - Material gratn In conformity with specifications yes no


If NO. attach gratn control graph for justification
Indicating method used with photographs of material tested
o o
J - Control of continuity index Icw

CONTROL OF CONTINUITY INDEX

APPLIANCE USED TRANSMITTER TYPE MEASUREMENT ACCURACY :


RECEIVER TYPE

Block number Face Oistance between


Emetteur & Receiver Ti"", Rate Jcw COtt1ENTS
(in cm) (In s) (In m/s) (visible cracking)

1 AA'
BB'
(CC' stratl)

4 - Form (dimensional coefficient)

Miscellaneous conrnents

ThiS sheet Is completed every day by the site manager responsible for rockfnl reception. If
any delivery is judged not to be in conformity with the speCifications, the geologist an the
approved laboratory are asked to carry out measurements, analyse the causes and advise the
Contractor an the appropriate measures to take.
607

Depending on laboratory analyses, quarries are registered as approved


sources for materials in classes A, B or C. An opinion is also expres-
sed as to the compatibility of the quarry with the quantities of rock-
fill required in normal operating conditions.

5 - INSPECTION AND CONTROL OF SUPPLIES

Controls are implemented on stocks close to the place in which they


will be used. They include:

- granularity control

The quarry operator or the transporter are responsible for the supply
and must respect the specifications. To facilitate controls and elimi-
nate errors, "front axis" scales exist which weigh every loaded block.
Material that is not in conformity with specifications must be refused
and withdrawn from stock.

A typical supply control checklist is shown below.

- continuity index control

Normally, supplies emanate from approved quarries only and thus should
be problem-free. Lodes do present anomalies, however, and it is sen-
sible to carry out controls when starting a site as well as batch sam-
pling and whenever so requested by the Contractor.

In general, controls are statistical. A visual inspection is made and


measurements are taken in suspect rocks. If checks show that one or two
blocks do not meet specifications, they must be rejected. If a high
percentage is rejected, the whole batch should be rejected and tests
carry out to determine the cause. This may be due to a local problem in
a quarry (accident, fault) or quarry operator error. Consequences are
determined with the Contractor. Shape is also tested at the same time
as wave propagation speed.

Contractors are responsible for deciding what technical or economic


measurements should be taken if supplies do not meet specifications.
This underline3 the importance of pre-purchasing analyses to confirm
that the quarries proposed by the firm ean supply. both ~he quantity and
the quality of rockfill required.

6 - PREPARATION CONTROL

This subject is also important, but only mentioned in passing. It is


vital to take special care with structural geometry. To do this. me-
thods include lateral echographic scanning of the immersed sections of
structures and beds using lateral sonar. Dyke walls can also be measu-
red to evaluate existing densities using diagraphic and liquid contai-
ner methods.
608

7 - CONCLUSION

Hydraulic engineers base most of the formulae used to scale rockfills


on rockfill mass and shape. Rock density is a scaling factor that pro-
ject designers can modulate to meet local availabilities and the posi-
tion of the block in the structure. This criterion does not, however,
guarantee gnod quality blocks and we propose specifications on the pro-
perties required to obtain this objective. Two types of properties have
been suggested

1 - Intrinsic properties inherent in the nature of the materials and


measured in the lode or quarry. These are basically the continuity
index (Ic), the degree of cracking (Df) and the Deval wet (Dh). Depen-
ding on the environment and the conditions specific to the structure,
additional properties can be added such as porosity to gauge a mate-
rial's sensitivity to frost ...

Rocks should be ranked in 3 classes according to their properties

- Classe Ai : good quality rockfill for use in high stress zones

Classe Bi good quality rockfill for use in medium stress


zones

- CIa sse Ci : rockfill for use in low stress zones

2 Properties of supplies including the conformity with Contractor


specifications for granularity, shape and quality as well the effect of
scaling on intrinsic properties.

Well up-stream from the site, Contractors should nevertheless evaluate


the true pos sibilities of local resource potential, particularly in
terms of the percentages of different types of blocks of different ca-
tegories that are available. Microsismic techniques have enabled more
accurate quantitative analyses to be made although they do not replace
conventional techniques.
609

BIBLIOGRAPHY

TOURENQ C, FOURMAINTRAUX 0, DENIS A. (1971) - Propagation


des ondes et discontinuite des roches, CR Symposium - Soc.
Meca-Roches, NANCY - Oct.

DENIS A, PANET M., TOURENQ C (1971) - L'identification des


roches par l'indice de continuite Congres Int. Meca-
Roches, MONTREUX

ALLARD P, BLANCHIER M. ( 1980 ) Estimation previsionnel1e


de production d' enrochements, Rapp. du C.E.T.E. d'AIX-en-
PROVENCE

ARNAULD, CHEVALIER (1984) Le bloc Acropole Travaux,


supp. Jan.

CHEVASSU G. (1979) - Enrochements, note sur l'aptitude des


principales roches exploi tees en Bretagne a la production
de gros enrochements de quali te Rapp. du LR de SAINT-
BRIEUC - Sept.

EVRARD H, MASSON M. (1977) Ressources en materiaux de


Basse-Normandie - Bull. Liaison Labo Ponts et Chaussees n
28 - Nov-Dec. pp. 2-1 to 2-12

VILLAIN J, MONNET J. (1980/1981) - Etude des enrochements


fluviaux Alpes du Nord, Jura Meridional - Rapp. du C.E.T.E.
de LYON

Laboratoire Regional de BLOIS, C.E.T.E. Normandie-Centre


(1985) - Methodologie de contr6le de l'etat des digues par
sonar lateral - Note technique

Laboratoire Central des Ponts et Chaussees EVRARD, ALLARD,


CHEVASSU, ROBERT, VILLAIN (1989) - Les Enrochements (page
107)
CHAPTER 20
ASPHALTIC CONCRETE FACE DAMS
J. L JUSTO

1. Introduction

Asphaltic concrete face rockfill dams are those in which this material is used as the
impervious element.
When asphaltic concrete is applied on the upstream face of an earth dam, it is not
clear if the object is imperviousness. as it might be slope protection.
According to our records, the first dam constructed with an asphaltic concrete
face was Mulungushi Dam (v. table I), made in Zambia in 1923.
The dam, 46 m high, was built with dumped friable micaschist. The attempt was a
failure: the dam settled 1.40 m and, as-a result, the facing broke along a line
parallel to the slope. The leakage was 7.1 m3 /s in 1925, and th e water was
emerging at the downstream face of the rockfill at heights of as much as 18 m. This
shows, also, that the rockfill was not free-draining.
The asphaltic facing was 100 mm thick at the crest, increasing 10 150 mrn below
level 1086 ID. It was formed by a mixture of gravel and a propietary brand of
asphalt. According to Legge (1970), perhaps it became embrittled by the high am-
bient and sun temperatures prevalent in that area before the rairu begin, and was tllttS
not capable of flexing under water load while subject to settlements, and failed.
Remedial measures consisted mainly in dumping impervious material into the
water and over the upstream face (fig. 1). The operation took four years (up to
1929), at the end of which some 115,000 m3 of material had been dumped and the
leaks staunched.
The water level had, by then, dropped sufficie ntly to expose the top of the clay
blanket. A reinforced concrete facing was constructed on it, which extended some
meter below the blanket.
After repair the dam has functioned satisfactorily without any signs of significant
leakage, for at least 40 years.

e 1991 K/,(...u Ac"demi.. P"hlis/rers.


'"
E. Maraulta das Ne\~$ (ed.). Ad,a uces;u Ro.:/.:jill SII"UClures. 611-650.
612

Table I. Performance of rockfill dams with asphaltic concrete facing (alphabetic order).
I I I I
I Rockfill construction IPostcons-1 I
Name, I Up. I Itructive I I
Country I Islopel I I I I I I cS IRevetment I I
& I I l:H I Rock IPlacement I I n I Eoedlsmaxl maxi structure I I Performance I I
year of I I I I I I I I I I I I I
Icompl et i on I H I type I method I I % I(MPa) (cm)l(cm)1 (em) I I I I
I I I I I I I I I I I I I
IAlesani 65 I 1.6 I I I I I 0.8 11O(b)+ I 3 INo cracks 110 1
I(France) I I I I I I 12x6( i) I ISome leakage I I
11969 I I I I I I I I lat joint I I
I I I I I I I I I lof gallery I I
I I I I I I I I I I I I
IAnchal 25 I 2 I I I I I 110(i) I 1.71 I 81
I(France) I I I I I I I I I I I
11986 I I I I I I I I I I I
I I I
I I I I I I I I I I
IAlmendra 30 I 1.751 gr. IVibrated I 1.71 I I 14( i )+6(d)+ I 0 I I 11
III y III I 29 I I I No I I I I 17( i) I I I I
I(Spain) I I I Isluicing I I I I I I I I I
11968 I I I I I I I I I I I I I
I I I I I I I I I I I I I
IBigge 56 I 1.6 IShale I c. : I I I 46 I I 14(l)+6(i)+1.005ISome blistersll01
I(Germany) I I and 13Mg platel 0.81 I I 6.3110.3111(d)+ 116 1
11964 I Iclayeyl5Mg vib. I I I I I 12x6(i)+ I I
I I Igray I ra. I 1.21 I I I 10.5(s) I I
I I Iwacke I I I I I I I I I
I I I I I I I I I I I
IBou 55 10.81 ISound I Hand I 126.51 9.5 1>6.31 11220(pc)+13(1)lpr.c. outer ll OI
IHanifia Is. & I & I I I I I 12x6(i)+ I revetment hasl121
I(Algeria) Icongl~1 derrick I I I I I I (p. r.c.) I suffered I I
11937 Imerate I I I I I I I progressive I I
I I I I I I I I I cracking & I I
I I I I I I I I Isl iding froml I
I I I I I 1 I I I 1938 ti II I I
I I I I I I I I 11970 (35 cm I I
I I I I I I I I 1 sl iding) I 1
I I 1 I I I 1 1 1 I
Icoo LowerlD30 1 2 IPhyl i-I I I I 11.5(i)+ I 1101
I(Belgium) IU21 1 1 tic 1 I 1 1 16(d)+6(i) I 1 1
11969 I I Ishale I I I I I I I I
I
I I I I I I I I I I I I
IDorlay I 44 I 1.7 I I I 1 I 112(b)+ 15-6 IExcellent 110 1
I(FranCe) I I I 1 1 I I 12x6(i) I I 1 I
11972 I I I I I 1 I I I I I I
I I I I I I I I I I I I I
613

Table I. Performance of rockfill dams with asphaltic concrete facing (alphabetic order)(continued).
, ! I I I
I Rockfill construction IPostcons1 I I
Name, I Up. I Itructive I I I
Country I IslopeI-I-----,-,---r-,-.,.----,-,--+I--r,--IIRevetment I I I
& I I 1:H I Rock IPlacementl n I EOedlsmaxlOmaxlstructure I I Performance I I
yea r of I I I I I I I I I I I I I
Icompletionl I H I type I method I % I(MPa)l(cm)l(cm)1 (em) I I I I
I I I I I I I I I I I I
IOungonnelll 17 I 1,7 I I I I I 17.5( i)+ I o I Some sea l 110 I
I(U.K.) I I I I I I I 112.5(d)+ Icoat ripples I I
11970 I I I I I I I 12x5 ( i ) I 10 11111 deep I I
I I I I I I I I 1 lafter prolan I I
I I I I I I I I I Iged drought I I
I I I I I I I I I I I I I
IEggberg 131.51 1.75lSlightl Rubber I 0.41 75 I 9 I 13(b)+4+ INa damage I 61
I(Germany) I I l l y ce1 tired I I I I 16(i) I 110 I
11964 I I Imented I r. I I I I I I 1121
I I I lsi l i cel I I I I I I I I
I I I lous s1 I I I I I I I
I I I I& whe~1 I I I I I I I
I I I Ithered I I I I I I I I
I I I I gn. I I I I I I I I
I I I I I I I I I I I I
IEl Siberia I 82 I 1.6 ISound I Vib. I 2 I I 5.41 15 IPl inth displ'!.l I
I (Spain) I I I trachyl r. I I I I Iced 6cm down. I I
11978 I I Isyeni1 I I I I 1& broken. Twol I
I I I I te I I I I I Icracks conti1 I
I I I I I I I I I Inued 1.8 mini I
I I I I I I I I I lasphalt. I I
I I I I I I I I I I I I
IGenkel I 43 I 2.251 Slate I c. : 10.45 1 I 1I 16(l)+ I 0.3lCracks near 1101
I (Germany) I I 1(2.5- 12.5Mg I I I I 12x3(i)+ I berm 1121
11952 I I 138 cm)lvib.ra. &1 I I I 112(d)+2(b)1 1301
I I I I I steel r1 I I I 1+3x3(i) I I I
I I I I I I I I I I I I I I
IGhrib I 65 0.71 ISound IHand & I 125.51 73 I 35 I 18(pc)+ I 25 I 1953:some r. 1101
I(Algeria) I l. I derrick I I I I 17 I 12x6(i)+ 10.1 Ie. plates I I
11935 I I I I in I 110(prc) I Isl ided down: I I
I I I I Ifou~ I I I( i) in good I I
I I I I Ida-I I I Icondi t i on. I I
I I I I I tionl I I 11963: Separa-I I
I I I I I I I I It i on of a. c. I I
I I I I I I I I Iplates; de I I
I I I I I I I I Isign for reo I I
I I I I I I I I Ibuilding I I
I I I [ I I I 1 I I I
614

Table I. Performance of rockfill dams with asphaltic concrete facing (alphabetic order)(continued).
I I I I I
I I I Rockfill construction Ipostcons-I
Name, I I Up. I Itruct ive I
Country I Islopel I I I I I IRevetment I
& I I l:H I Rock IPlacement I n I Eoedl smax I <5maxi structure I I Performance I
year of I I I I I I I I I I I I
ICDq)letionl I H I type I method I X I (MPa) I(cm) I (em) I (em) I I I
I I I I I I I I I I I I I
IGrossee I 57 I 1.5 I gn. I I I I I 18(b)+8( i) I IFailure & de-I
I(Austria) I I I I I I I I I I Ipressions. Del
11980 I I I I I I I I I I Ifects in con=-I
I I I I I I I I I I I Itact wi th cut I
I I I I I I I I I I I-off. C"",,le-I
I I I I I I I I I I Ite repair of I
I I I I I I I I I I I lower portionl
I I I I I I I I I I Inecessa ry 63% I
I I I I I I I I I I I I
IGuajaraz 48 I 1. 75 1 gr. 110Mg vib10.3- Y =1 I I 15(l)+3(b)+1 0.51 Good I
I(Spain) I I r. I 1 19.61 I I IZx6(i) I I I
11972 I I I I I I I I I I
I I I I I I I I I I I I I
IHenne 58 I 2.251sound Ic.: 10.8 I 180 I 2.41 4.516(1)+ 10- 21 Excellent 110 1
I(Germany) Il. & 13Mg platel I I I 12x3.5(i)+ I Slight pea- 1121
11955 Igray- I & I I I I 11O(d)+ Iling of mas- I I
I I Iwacke 110Mg vi - I I I I 13x3(i)+ I tic I I
I I I I brator I I I I 12xO.5(s) I I I
I I I I I I I I I I I I
IHochwurtenl 55 I 1.65lGravell I I I I 18(b)+8( i) IContact with 1291
I(Austri a) I I lmorraij I I I I I I cut-off I I
11980 I I lne &-1 I I I I I Itrench dama- I I
I I I I rock I I I I I I Iged, depres- I I
I I I I I I I I I I I sions & I I
I I I I I I I I I I Icracks. New I I
I I I I I I I I I I I impervious f~1 I
I I I I I I I I I I Icing in 75%. I I
I I I I I I I I I I I I I I
IHuesna I 71 I 1.6 ISilic~1 Vib. r. I 0.91 Y d=1 35 I 1 I 16(b)+8(i) 0.3lExcellent 1291
I(Spain) I I I ous I 10 Mg I 120 .61 I I I I I I
\1990 I I Islate 18 passes I I I I I I I I I
I I I I I I I I I I I I I I
\I rit Emda I 75 I 1.61 l. Ie. with 10 25 1 I 49- I 6.41 1O.8j12(pc)+ 5 ISatisfactory I 81
I(Algeria) I I I Ir. & vibl-l I I 196 I I 12x6(i)+ IProblems 110 1
11954 I I I Isledge 1-1 I I I I le15(rc) Iaccording tol121
I I I I 1-2 passes I I I I I I I (1) 125 1
I I I I I sluiced I I I I I I I I I
I I I I I I I I I I I I I I
615

Table I. Performance of rockfill dams with asphaltic concrete facing (alphabetic order)(continued).
I I I I
I Rockf ill construction IPostcons1 I
Name, I Up. I Itructille I I
Country I Islopel I I I I I IRevetment I I
& I I l:H I Rock IPlacementl n I EOedlsmaxlOmaxlstructure I I Performance I I
year of I I I I I I I I I I I I I
Icompletionl I H I type I method I % I(MPa)l(cm)l(cm)1 (cm) I I I I
I I I I I I I I I I I I I I
IKonoyama 33 I 1.8 I I I I I I 18(m)+8(1)+1 o 10penings at 1101
I(Japan) I I I I I I I 15(b)+ I I joints. Good I I
11971 I I I I I I I 11O' 15(i) I Icondition af1 I
I I I I I I I I I I Iter repairingl I
I I I I I I I I I I I I I
ILicheyu 57 I 1.75IAnde IOirectio1 36 I I 130.81 lOry mason' I yeslCracks at 1321
I (China) I I site Inal blas1 I I I I Iry 100 I Ijoints with I I
11977 I I I ting I I I I I 12x5(b)+ I Icutoff in I I
I I I I I I I I I 12x5(i) I 136.6 m. Lar' I I
I I I I I I I I I I I Igest relatillel I
I I I I I I I I I I I Ideflect i on I I
I I I I I I I I I I I I 15 cm I I
I I I I I I I I I I I I I I
IMartin I 55 I 1.5 ISlate IVib. r. I 0.81 Yd=1 I 8.51 12x6(l)+ I INo damage I I
IGonzalo I I I & 114 Mg I 120.11 I I 16(b)+8( i ) I I I I
I(Spain) I I Igray' I 6 passes I I I I I I I I
11989 I I Iwacke I I I I I I I I I
I I I I I I I I I I I I I
IMiyama 175 .5 1 1. 9 IRhyolil I I 10 I 13.5 (1)+ I 3.6l Good 91
I(Japan) I I Ite & I I I I 16(i)+6(d)+1 I I
11973 I I landesj...l I I I 12x6(i) I I I
I I I Itic I I I I I I I I
I I I Ituff I I I I I I I I
I I I Ibrec- I I I I I I I
I I I I cia I I I I I I I
I I I I I I I I I I I I
IMontgomeryl34.51 1.7 ISound d. I 10 I I 9I 12.5- 7.5(1) I o ISome joints 1221
I(USA) I I I gr. slui ced Imax. I I I 1+10+9+ I lopened near I I
11957 I I I I I I I 17.5(i) I Ithe crest I I
I I I I I I I I I I I I I
IMulungush i I 46 2 IFria- d. I I 301 1421 110-15(i) 171001ln 1925 appeal171
I(Zambia) I I ble I I 10 I I I I Ired a crack I I
11923 I Imica- I I I I I I Iparalled to I I
I I Ichist I I I I I I Islope, wider I I
I I I I I I I I I lat crest I I
I I I I I I I I I I I I
INegratin I 75 1.6 I IC. by 4 I 0.91 I 32.81 241 15 (l )+6( i )+ I o INo damage I 11
I(Spain) I I Ipasses ofl I I I I 19(d)+ I I I 41
11985 I I Illib. r. I I I I I 12x5(i) I I I I
I I I I I I I I I I I I I I
616

Table I. Performance of rockfill dams with asphaltic concrete facing (alphabetic order)(continued).
I I I I I
I I Rockfill construction IPostcons'l I
Name, I Up. I Itructive 1 1
Country I Islopel I I I I I IRevetment 1 1
& I I 1:H 1 Rock IPlacement I n I EoedlsmaxlOmaxlstructure 1 I Performance I I
year of 1 I I 1 1 1 1 1 1 1 I 1 I
Icoq:>letionl 1 H 1 type 1 method I X I(MPa)l(cm)l(cm)1 (cm) I I I I
I I I I 1 I I I 1 1 1 I 1 I I
IOgliastro 122.2122.4IRippedl C. 6 1 0.81 Y= 1 1 2I 18(d)+6( i) I 7.31 1121
I(ltaly) I I Iporous 1 passes I 118.61 Ifounl I I 2.51 I I
11970 I I Icalca1 vib.r. I I I Ida, I I I I I
I I I Irenitel I I I Itionl I I I 1 I
I I I I I I I I I I I I I I 1
IOtsumata 1 52 1 1.7 1 1 I I I I I 17(l)+5(i)+1 I Good condi' 1101
1(Japan) 1 I 18(d)+ I I tion
11968
I 1 I I I
I , I
I I
12x5(i) , I I
, I I
I , 1
I
1
1 I I I
I
,I
I I I
I
I
I I
I I
IPerlenbachl 18 I 1.751 gray I I I I I I 16(l)+ I IRenewal of s I I
I(Germany) I I I wacke I I I I I I 12.5(b)+ I land joints inl I
11954 I I I I I I I I I 12x3(i) I I 1985 1 I
I I I I I I I I I I I I I I I
IPrims I 62 I 1.75ISlate,1 C. by I I y= I 20 I 50 I I I 10 I No damage 13 11
I(Germany) I I I some I rubber I 121.51 I I I I 0.11 I I
11978 I I Iquart'ltired r1 I I I I I I I I I
, I I I zite I I I I I I I I I
I I I I I I I I I I I I I
IRadoina I 42 I 0.8 I I d. well 10'1 11.851 115(pc)+ I 2'31 No damage 110 1
I(Yugosla' I I I I slui ced 15 I I I 12x4.5(i)+ I I 1121
IVia) 1959 I I I I I I I 112(c) I I I I
I I I I I I I I I I I 1 I
IRiveris I 45 2 I I I I I 16(l)+4(i)+1 I Good concJi 110 I
I (Germany) , , I 18(d)+6(b)+1 I tion
1 1955 , I I I
12x4( i) , 1 I
I
I
I I
I
I
I
I 1
1 I
, I I I I
I I I
1 I
ISabigawa 190.5 2.0 ISili- 1 IYd= I 1 1 14(m)+4(l )+1 I 1111
I(Japan) I Iceous I I 19 I I I 16( i )"8(d)" I I I I
I I I tuff I I I I I 13x5( i) I I I
I I porphyl I I 1 I I I I I
I I rite-I I I I I I I I I
I
ISainte
I
42 1.7
I
I
I I
1
I
I 6.3
I I
,10(b)+ , Very good 1I 81I
I
ICeci le , I I 12x6( i) I behaviour 11 01
Id'Andorge I I I I I Saggings in I I
I (France) I I I I I seal coat I I
11967 I I I I I I I
I I I I I I I I I
ISalagou 63 I 1.5 I I 180 I 11 I o I No damage I 81
I(France) I I I I I I I I
11971 I I I I I I I I
I I I I I I I I I
617

Table I. Performance of rockfill dams with asphaltic concrete facing (alphabetic order)(continued).
I I I I
I Rockfill construction IPos tcons-I I
Name, I Up. Itructive I I
Country I Islopel I I I I I IRevetment I I
& I I l:H I Rock IPlacement I n I EOedlsmaxlOmaxlstructure I I Performance I I
year of I I I I I I I I I I I I I
Icomplet i on I I H I type I method I % I(MPa)l(em)l(em)1 (em) I I I I
I I I I I I I I I I I I I I I
ISantillanal 40 I 1.751Sound ISMg vib_ 10 . 5-1 I I I 14.5(m)+ 1<1.91 Elephant's I 11
I II I I I gr_ I r. 4-5 1.3 I I I I 13.5(b)+ I x I skin. Some I I
I(Spain) I I I I passes I I I I 16(i)+9(d)+ll0- 21 flow. Exce- 1 I
11969 I I I I sluiced I I I I 110(i) I Illent result I I
I I I I I I I I I I I I I I
IScotts I 46 I 1.7 I Soft I C. 4 0.91 Y= I 34 138.81 I Below a I 125 ISubhorizontalI 71
IPeak I I I argi-I passes 120.61 I I I depth of I (6)lerack at con-I I
I(Austra- I I III ite 110Mg vib. I I I I I 30 m: I Itaet of gra- I I
Ilia) 1972 I I I r. I I I I I 3 layers I Ivel toe and I I
I I I I I I I I I 18.5 cm I Irock fill, and I I
I I I I I I I I I Above: I Ihole 1 mlongl I
I I I I I I I I I 2 layers I 10.5 mwide. I I
I I I I I I I I I 14 cm I IInternal ero-I I
I I I I I I I I I I Ision of gra- I I
I I I I I I I I I I I vel toe. I I
I I I I I I I I I I I I I
IShibianyu I 85 I 1.8 I gn. IOirectio-157.31 I 138.41 lOry mason-I yesl bl i sters, I I
I(China) I I I I nal I I I I I Iry 150-200 1 I caves & I I
11978 I I I Iblasting I I I I I 112(b)+ I I eraks I I
I I I I I I I I 12x5(i) up I I I I
I I I I I I I I 112(b)+ I I I I
I I I I I I I I 13x7( i )down I I I I
I I I I I I I I I I I I I
ISteinbach I 40 I 1.75IWeath~1 Rubber- I I 36 I 31 I 15(l)+4(i)+1 I No damage 1121
I(Germany) I I I red I tired I I I I 17(d)+ I I I I
11963 I I Ischistl r. I I I I 12x4(i) I I I I
I I I I with I I I I I I I I I I
I I I IgraveL! I I I I I I I I I
I I I I size I I I I I I I I I I
I I I I I I I I I I I I I I
ITataragi 164 . 51 1.8 I gr. I IY = I 53.61 9.61 18(l)+5(i)+1 o I Good I 31
I(Italy) I I I I 119 .61 I I 18(d)+ I I I I
11973 I I I I I I I I 12x6(i) I I I I
I I I I I I I I I I I I I I
IVallon Doll 45 I 2 I l_ I I I I I 110(d)+ I 1.5lLocal flow ofl 81
I(France) I I I I I I I I 12x6(i) I (O.2~6 cm downslo-ll01
11973 I I I I I I I I I I lpe of a.c. I I
I I I I I I I I I I I I I I
618

Table I. Performance of rockfill dams with asphaltic concrete facing (alphabetic order)(continued).
I I
I Rockfill construction IPostcons-1 I I
Name, I Up. I Itructive I I I
Country Islopel I I I I I IRevetment I I
& I I:H Rock IPlacement I n I Eoedl smaxlOmaxl structure I P..:rformance I
year of I I I I I I I I
Ic~letionl I type I method I % I(MPa) I(cm) I (em) I (cm) I I
I I I I I I I I I I I I I
IVenema I 51 I 1.7 gn. IDown. d. I :5201 I 29.4 11.21 12 110( l)+ I 4.51 No damage 11 0 1
I(Norway) I I Up. BMg I 1.51 23-1 I 13x6( i) I I 1121
11963 I I vib. r. I I 35 I I I I I I I
I I I (10-15 I I I I I I I I I
I I I passes) I I I I I I I I I
I I I I I I I I I I I I
IlIahnbach 145.61 1.6 C. by 1<1. 51 I 1-21 16(l)+4(i)+1 INo repairs orl121
I(Germany) I I rammer & I I I I I l1(d)+9( i) I I maintenance 125 1
11957 I I 10 Mg I I I I I I I I I
I I I vibratorl I I I I I I I I
I I I I I I I I I I I I
Illinscar I 53 I 1.7 s. Vib. r. I 1.71 22 I 20 12 119 . 5 1 2 layers I 37 I Cracks at 120 1
I(U.K.) I I 13.5 Mg I I I I I I I union of I I
11974 I I 14 passes I I I I I I I facing with I I
I I I I I I I I I I I culvert I I
a.c. = asphaltic concrete (m) = macadam
(b) = binding layer (p.c.) = porous concrete
c. = compacted (p.r.c.) = porous reinforced concrete
d = dumped r. = roller
(d) = drainage layer ra = ranmer
down. = downstream (r.c.) = reinforced concrete
found. = foundation s. = sandstone
gn_ = gnei ss (s) = sealing coat
gr. = granite smax = maximum crest settlement
(i) = impervious layer up. = upstream
l. = limestone vib. = vibratory
(l) = levelling layer 0max = maximum displacement normal to facing
Y and Yd in KN/m3
619

1108
v

Borehole

Sat uration line


1060
---------=~==-------
v
o 10 20 30 40 50 60 M
1~~---e~~1--~~~1--41
Sea I e

(j) Aditional rockfill


30 em Concrete facing
Dumped clayey soil

Figure 1. Mulungushi Dam section after repairs (Legge, 1970).

The second important attempt is formed by a series of high dams (Ghrib, Bou
Hanifia, Radoina), with rather steep upstream facing (mainly 1:0.7 to 1:0.8). The
first of them, Ghrib dam, 65 m high and constructed in 1935, is a landmark in the
history of dam construction. The impervious facing consisted in a layer of asphaltic
concrete 9 to 12 cm thick, supported by a layer of porous concrete 12 to 20 cm thick
and covered by reinforced porous concrete slabs as a defense against sun rays and
rock fall.
In most of these dams the concrete slabs have progressively slided down and have
finally been taken off.
Ghrib and Bou Hanifia were made with derrick and hand placed rockfill, and
Radoina with well sluiced dumped rock.
Iril Emda, 75 m high, was constructed in 1954 with similar facing, but with
compacted rockfill, a slope less steep (1: 1.6) and non-porous reinforced concrete
protection.
According to our records, the last facing of this type corresponds to an earth dam:
Kruth Wildenstein, 37 m high, constructed in 1964. From the first partial impoun-
ding the facing suffered local ruptures, and the leakage was 65 lis. The reinforced
concrete slabs were substituted in 1974 by two layers of asphaltic concrete.
According to the French Comunittee on Large Dams, it is necessary to forbid the
association asphaltic concrete-reinforced concrete owing to the lack of adherence on
the slope and to the different thermal behaviour of both materials.
620

They add that no example of good behaviour in the long lUn of this association exists
(Ghrib 1936, Bou Hanifia, 1938, lril-Emda, 1954).
Notwithstanding these asserts, which are true as far as the reinforced porous
concrete cover is concerned, the leakage of all these dams has been quite small (v.
table J) and the behaviour of the asphaltic concrete quite good.
The next step is given by the Federal Republic of Germany, which starts, in 1952,
with Genkel rockfill dam, 43 m high, a series of dams of well compacted rockfill
with an asphaltic concrete face (type A) formed by a sandwich stlUcture (fig. 2)
consisting of a bituminous concrete drainage layer included between two dense
bituminous concrete layers (the external one generally placed in at least two
courses), all that placed on a bituminous levelling and binding layer. The reason for
this particular structure is seepage collection and measurement in a toe drainage
gallery (fig. 3).

Mastic coating spread in /lenhnli,, concrete compacted in 3 layers


two coats (0.5 cm thick) Asphalt binder
Drain of coarse grovel lightly bound with asphalt
Itic concrete compacted in 2 layers
rmmJ~1Jj~"""'-L..cY'.IIng course of asphalt bound crushed rock

Compacted rock embankment (surface


voids filled with leon concrete)

Drainage gallery
(see detai I)

Rock line ( 0)

Figure 2. Genkel dam. Typical cross-section with details of facing (Sherard et al.,
1963).

The series is continued by Henne, Riveris, Wahnbach, Steinbach, Bigge, Glems


and Nagold (1967). The steepest upstream slope corresponds to Wahnbach (1:1.6)
and the flattest to Genkel (1:2.25).
621

The structure of the facing is similar in all of them (fig. 2): tackcoat, 4-7 cm
binding-levelling layer, 4-7 em secondary impervious layer, 6-12 em drainage layer,
0-6 em binder course, 7-12 cm primary impervious layer and seal coat.

,L ,L
1. 50
(b)
Dimensions in meters

Figure 3. Details of drainage gallery at upstream toe of Genkel dam (Sherard et a1.,
1963).

There is no tackcoat in Wahnbach, Steinbach and Bigge.


The heights range from 21 m (Glems) to 58 m (Henne).
Leakage in all these dams has been negligible and damage scarce.
Also the Germans start around 1954 a second type of asphaltic concrete face (type
B), in which a binder course, followed by a dense bituminous concrete impervious
layer, is superimposed on a levelling layer.
The series is initiated with Perlenbach dam, followed by Kessenhamm, Ulmbach,
Eggberg, R6nkhausen, Wehra, Weilerbad, Langenprozelten upper reservoir and
Kronenburg (1975).
In the three former dams the impervious layer is formed by two courses with
staggered joints. From Eggberg Dam (1965) only one course is placed.
622

The upstream slopes range from 1:1.75 (Perlenbach, Eggberg, Wehra) to 1:2
(Kessenhamm, and Langenprozelten). The heights range from 18 m (Perlenbach
and Kessenhamm) to 40 m (Wehra).
The structure of the facing is similar in all of them: tackcoat, levelling course 0-6
em, binder course 0-8 em, impervious course 5-10.5 em, and seal coat. There is no
tackcoat in Kessenhamm, Ulmbach, Weilerbad and Langenprozelten.
We only know exactly the behaviour of Eggberg dam, which has suffered no
damage. At Perlenbach renewal of mastic and regeneration of joints was made 31
years after the end of construction (v. Strabag).
The fact is that, after 1967, type B seems to have superseded type A in the Federal
Republic of Germany (v. !Cold, 1982).
Montgomery dam, 34.5 m high, constructed in 1957 in the U.S.A., is the first
successful dam made of dumped rockfill. This dam is, according to our records, the
first dam to incorporate a slab type plinth (fig. 4). In this type of plinth the
displacement gradients of the face are quite small. It has a type B structure, but
with a very thick impervious layer (three courses with a whole thickness of 26.5
em).

ORIGINAL GROUND SURFACE 30 em MIN. RADIUS

SURFACES THOROUGHLY CLEANED AND


SPECIALLY TREATED
1"- 0.40 0.15 0.15
~~~~~~~~

ROCK SURFACE
:'-.'-... ~.

1 MIN. 2.5 MAX. - - - - - i THEORETICAL HEEL LINE

J
OF DAM_-+- 0.30 OPEN DRAIN
REINF. ST EEL ARIABLE
L...,lj:::~t--6cm GROUT PIPES EMBEDDED IN CUT-OFF
POUR AGAINST UNDISTURBED :' AT APROX. 20 em
FACE OF EXCAVATION .45
II
O.91IMIN.
GROUT HO L E S -~--~.~ I
-->-L
DIMENSIONS IN METERS

Figure 4. Plinth at Montgomery dam (Scheidenhelm and Snethlage, 1960).


623

In 1977 and 1978 asphaltic concrete facings are applied on two Chinese rockfill
dams whose lower 65% height is constructed by directional blasting (v. table I).
Shibianyu dam is 85 m high. Post-constructive settlement was important 38.4 cm,
and both dams suffered important damage and leakage that seems to have been
repaired successfully.
The height record for asphaltic concrete face rockfill dams corresponds, according
to our records, to Sabigawa dam, 90. 5 m high, constructed in Japan around 1988.
There is, at present, a trend towards the use of steeper upstream slopes, up to
1:1.5. Examples of that are, in Spain, Martin Gonzalo dam (1990), 55 m high, in
Austria Grossee (1980), 57 m high, and in France Salagou (1971), 63 m high.
Blisters have appeared in many asphaltic facings (v. table I). It seems that they are
produced by drops of water trapped between two impervious courses. In order to
avoid these blisters, construction methods have been developped to place linings
with sufficient thickness (up to 10-12 cm) in only one layer (v. Haas et aI., 1988).
This is the trend in many of the lattest asphaltic concrete facings (v. Icold, 1982). In
the F. R. of Germany, Austria and Switzerland, the dams built after 1971 follow
this trend.

2. Revetment structure

As indicated in the Introduction there are two types of facings used today, that we
have called types A and B. We shall describe them starting from the layer in
contact with the facing: in modern dams stones of small size (transition layer).
The sandwich type structure (type A) has been applied, according to our records,
in at least 23 rockfill dams and reservoirs, and consists, in the most complete case,
of the following layers (fig. 2):
1. A priming treatment with bituminous binder or cement of the subgrade to ensure
adhesion of the facing to the transition layer of the dam, and to provide a working
surface suitable to support the equipment for placing and compacting the next
layer. It is called tackcoat or stabilization.
2. A levelling layer made of coated chippings or bituminous mix, 3 to 8 cm thick. It
has also the function of binding between the subgrade and next layer.
3. A secondary impervious layer of dense asphaltic concrete, 4 to 7 cm thick.
4. A drainage layer, generally of open graded bistuminous mix, 6 to 12 cm thick, or
more if it is non-bituminous (Futaba rockfill dam in Japan, 22 cm).
5. A binding layer of bituminous concrete or coated chippings. When existing it has
from 2 to 6 cm.
6. A primary impervious layer of dense bituminous concrete, 6 to 12 cm thick.
7. A hot-applied mastic seal coat (2 to 4 mm), used for sealing the surface pores.
Such a seal coat prevents changes in the properties of bitumen by air oxidation
624

and by ultra-violet radiation as well as ice attack (v. SchOnian, 1989).


Japan continues using this type of structure for seismic reasons.
Type B structure may have the same structure as type A, except the secondary
impervious layer, with the following particularities:
1. The open asphalt levelling course may be, when it exists, 3 to 10 cm thick.
2. The drainage layer only exists in a few rockfill dams (Vallon Doll, Legadadi,
Valea de Pesti, Sainte Cecile D'Andorge).
3. The pervious asphalt binder course is from 2.5 to 12 cm thick, and may not exist
(Guajaraz dam in Spain, Valea de Pesti in Rumania and R6nkhausen in Germany).
4. Either the levelling or the binder course always exist.
5. The impervious layer should be from 7 to 12 cm thick. It should be placed in one
layer. The joints between adjacent finisher lanes should be gently reheated and
further compacted by heated tampers. Such welding results in total watertightness.
In some cases the type B structure may be extremely simple (fig. 5).

Sea I coot
------1
Impervious layer
---Bind~

Bituminous emulsion i
I

Support gravel &stone


(15-125mm)

Dimensions in m

Figure 5. Asphaltic facing at Huesna rockfill dam (Spain).

Indications about grading, bitumen properties and content of the different layers
are given by ICOLD (1982).
Bitumen content of the impervious layer usually ranges from 7.5 to 8.5%, and its
density ranges from 2140 to 2550 k~/m3. Maximum k value should be 10-7 cm/s,
the usual range being 10- 8 to 10- cm/s. According to Haas et al. (1988), the
maximum permissible porosity should be 3%. Kjaernsli et aI. (1966) arrive to the
same conclusion, as a means of assuring a k::: 10- 8 cm/ s. Another reason to
maintain porosity below 3% is to avoid oxidation (v. Sch6nian, 1989). Tschernutter
(1988) recommends less than 2%. Permeability usually decreases with time, which
625

is called "self-healing" (v. Kjaernsli et aI., 1966).


For the drainage layer k is usually::: 10-2 cmls (v. !cold, 1982).

3. Construction

Most of the asphaltic concrete facings have been made once the embankment was
ended.
Notwithstanding, in case of long slopes (above 100 m) or when summer construc-
tion periods in cold zones are short, asphalt concrete linnigs can without much
difficulty, be built in stages. The technique of jointing the linings at these stages is
easitly mastered by experienced cons tractors (fig. 6).

Figure 6. Martin Gonzalo rockfill dam repair. Work at contact between 1st and 2nd
phases of face construction (courtesy of Strabag).
626

Facing construction in two phases is also carried out when the capacity of the
bottom outlet is small and safety agains floods is sought, or, when the runoff is
small and it is important to advance impounding.
All this may apply specially to repairs (fig. 6).
The placing system by bands paraUed to the crest is receiving increasing attention,
specially when the length of the crest is greater than the height. In reservoirs this
technique greatly diminishes the time needed to reset the placing and finishing
machine on the completion of each band.
Notwithstanding the common method of placement is by bands along the slope
(figures 7 and 8).

WINCH POR,AL

SUPPLY CART ---,,=~


c:~~
__,-
i- ],&0-1 I
PAVER
i-- &1.--<

//~/// TRVC~ I
FO R TK ANSPCRT
VIBRATORY OF ASPHALT MIX
IN INSUlA1EO
ROLLER BUCKErs
2,1 m l

V I BRATQRY ROLLERS

~ J 2 To .
0----- -- ------ ------- -------------

AUXILIARY
WINCH
TRUCK

Figure 7. Paving equipment (Schewe and Geiseler, 1985).


627

Figure 8. Binder placement during 1st phase of Martin Gonzalo asphalt face
construction (courtesy of Strabag).
For reference compare with figure 6.

We follow the description by Schewe and Geiseler (1985):


The system is similar to the one of road paving, however modified for use on
slopes: winches operate on the dam crest and pull the paver and the rollers.
The paver operates on the maximum slope line, placing lanes 3.5 to 4.5 m wide,
jointless from bottom to top (fig. 9). The rollers follow, either behind the paver and
towed by a winch mounted at the rear of the paver, or on the previously laid lanes,
towed by winches located on the dam crest.
The hopper of the paver is usually not capable of taking the whole quantity of mix
required for a complete lane; therefore intermediate feeding is necessary.
A separate, winch-towed, supply cart runs up and down the slope and feeds the
paver. As it is not possible to dump the hot mix directly from the lorries into the
supply cart, insulated buckets, filled at the hot mix storage silo of the batch plant,
are emptied into the supply cart by means of a mobile crane standing on the dam
crest. Figure 7 gives a schematic view of this assembly of slope paving equipment,
and figure 8 shows a photograph of it.
628

Figure 9. Placing of impervious layer (2nd phase) in Martin Gonzalo repair


(courtesy of Strabag).

The winches are all mounted on a self-propelled winch portal, moving on tracks at
the crest. When the paver has finished a lane, it moves into a portal-like opening,
and the complete set of machinery moves into position for the next lane.
Two or three vibratory rollers are used for compaction.
On the freshly laid material acts first a light vibratory roller, of 1 to 1.5 Mg, towed
by a winch mounted on the rear end of the paver, or towed by a winch mounted on
the winch portal and operating on the previously laid lane.
A 4 to 5 Mg vibratory roller does the final compaction and smoothening of the
surface. This roller and its auxiliary winch truck operate independently of the main
unit, two or three lanes behind the paver.
The joints between the lanes are in general water-tight if the fresh material is laid
and compacted adjacent to the previous lane, while this one is still hot. However it
is advisable to give a special treatment to all the joints of the top impervious dense
asphaltic concrete. The asphaltic concrete in the joint area is reheated by quite long
joint heaters, which are slowly pulled from toe to top, followed by or combined with
hammering the reheated material with a tampering device, which welds the two
lanes.
Finally, by means of a squeegee on wheels towed by a winch, the hot mastic
629

asphalt surface protection is spread on the slope in lanes from toe to top.

4. The deform ability of asphaltic concrete related to the strains suffered by the
facing

A summary of this subject has been presented by Justo (1968 and 1973).
Asphaltic concrete is a visco-elastic material whose mechanical properties are
largely depended upon strain rate and temperature.
Figure 10 shows the dependence of tensile failure strain upon method of loading
and temperature.

10

8
~
~
c: 7

.....
0

If) 6
0'1
c:
.:>
0
5
0
L
0
L 4
0

...
C1l
::;J 3
0
LL 2

o ~ __ __
~ -L~ _ _ _ _ _ _- L_ _ _ _ _ _ _ _ ~

-10 o 5 20 40
Temperature (Oe)

Figure 10. Failure strain of asphaltic concrete as a function of method of loading


and temperature (Daicho, 1988).
1. Beam bending test. 4. Uniaxial tension test.
2. Flexibility test. 5. Upper limit of strain on crack ocurrence.
3. Biaxial tension test. 6. Lower limit of strain on crack occurrence.
Stress rate 10-30 kPa/min.
630

For temperatures larger than 5C the strain at failure is larger than 3%.
We have also indicated in that figure the upper and lower limits for crack
occurrence. Asphaltic concrete even fissured remains impervious up to failure.
Figure 11 shows the dependence of bending tensile strain upon strain rate and
temi,erature. Strain rates in the field are much smaller than 10-5 Is (10- 10 Is to
10- Is) and so tensile strains will be, for temperatures::: 5C, much larger tan 5%.

16 ~-----'----Tr-r------.------.

14

C1) 6
en
c
C1)
r- 4

Figure 11. Relationship between bending yield tensile strain, temperature and
strain rate (Ishii and Kamijo, 1988).

During permeability tests, discs of asphaltic concrete have been subject to bending
tensile strains of 25% at 20C and 22% at 5C without any leakage. These strains
are from two to four times the yield strains of figure 6, which are, so, safe.
Other authors (v. Justo, 1968) quote strains of 1 to 2.1 % before fissuring, and 1.35
to 6.3% for failure, with temperatures from 0 to 30, and strain rates larger than
5xlO- 4 /min. On the other han, for temperatures as low as -79C, the failure strain
might range from 1 to 0.2%.
Up to now, no dam has suffered damage due to excessive compression. The
reason may lie in the high failure strains of asphaltic concrete in compression.
Probably the higher tensile strains in a facing will appear at the junction of this
facing with the abutments.
631

Figure 12 shows the displacements and tensile strains of the asphaltic concrete
facing calculated, at the maximum section of Tataragi Dam, by the FE method. It
may be seen that the maximum tensile strain appears near the junction with the
foot gallery, and is near 1.1 %. Similar maxima tensile strains have been found by
Justo (1968 & 1973). In any case, they might be larger in the upper part of the
abutments or with more compressible rockfills. One proof of that is the cracking of

--I
the concrete membrane at Mulungushi dam (v. table I).

EL.185

180+_ __
E
c
o
i\
~ \
o
> 175
Q.)

170

@
o 1020 em
~

Figure 12. Displacements and strains of asphaltic concrete facing, along the maxi-
mum slope line, and in the vicinity of the inspection gallery of Tataragi Dam (Dai-
cho, 1988).
1. Displacements. a) Scale of displacements.
2. Tensile strain.
3. Tensile + bending strain.

In any case, the actual rate of increase of reservoir water pressure was, in Tataragi
dam, 10xlO- 4 kPa/min., 10-4 times smaller than at the laboratory (v. caption of
figure 5), and so the corresponding failure or cracking strains would be much
larger.
632

Figure 13 shows similar results for Sabigawa dam. We see that, anew, the largest
strains appear near the plinth: 1.1 % tensile strain, 3% compressive strain and 4.1 %
shear strain.

Figure 13. Strain distribution of facing in Sabigawa dam (Ishii and Karnijo, 1988).
1. Tensile strain.
2. Compressive strain.
3. Shear strain.

Shear strain perpendicular to the facing may be important when its support is not
homogeneous.
Thin plates, whatever flexible they are, will fissure if its support gets appreciably
fissured, specially if hoth lips of the crack do not remain in the same plane. On the
633

other hand, in a thick layer, the fissure started with the support may, thanks to the
plasticity of the material, dampen in its thickness without reaching the outer wall.
In this way the membrane will remain impervious.
The contact with the plinth, specially if the abutment is steep, and hydraulic
concrete structures should be specially treated (v. Delgado, 1988; Justo, 1973).
So as to ensure homogeneity, a layer of well compacted, small dimension material
should be placed under the facing.
A very favourable quality of asphaltic concrete is self-heating of initial fissures.
Wide temperature changes, intense frost, light, wind and waves may contribute to
the appearance of hair cracks in asphaltic concrete.

5. Plinths

The plinth is the element that connects the facing with the foundation.
As indicated in figures 8 and 9, the largest strains of the facing appear near the
plinth, due to the difference of rigidity of the foundation and embankment mate-
rial.
It is necessary to provide a connection at which the displacements gradients are
small.
The plinth of Montgomery dam provides quite small displacement gradients (fig.
4).
Notwithstanding, specially with sandwich type revetment, often a drainage gallery
is provided (fig. 3) to collect leakage through the facing and control it. This gallery
allows also regrouting of the cut-off without emptying the reservoir. Usually this
drainage gallery emerges largelly above the foundation to decrease excavation in
rock. The problem is enlarged when the plinth is used as a cofferdam in the river
bed.
If the foundation of the plinth requires deeper excavation than expected in design,
the plinth should be recalculated and, in any case, the excess concrete at the base
should be joined by reinforcement to the body ofthe plinth. Otherwise water
pressure could produce a rupture at the plinth (fig. 14).
The drainage gallery also produces large displacement gradients at the joint. At
least a curvature should be given at the connection with the facing (fig. 3) to
decrease these gradients and, consequently, strains at the facing.
The design of figure 15 provides very small displacement gradients and strains.
634

o 20 40 60 80 100 m
F- - I

@0.30
0.20 ~L 2.50 Ll L 0.70

II 1l'
-t-
1.50

1.00
,I, .... ,
CD\ 2.00

. 'i
. .. 1 3 .00

=-n"""~=rT6.TT!00::P"m="J"'t1 l~o~ f-
Figure 14. Failure of EI Siberio plinth (Alonso Franco and Soriano, 1988).
A. Schematic view of plinth. j) crack construction joint.
B. Failure at area "b". s) = detachment of plinth from
C. Failure at area "c". foundation.
D. Failure at area "d". t) rupture at oblique angle.

If the drainage gallery is not considered necessary, a slab-type plinth provides


quite small displacement gradients (fig. 16).
Figure 17 shows the connection of Martin Gonzalo plinth (cofferdam) with the
asphalt face.
635

Sandwic h tipe
aspha Itic fa cing
\

We II graded crushed
I_~o_ne (1 - 7 c~ __ _
3.001

1 \ Grout curtain Piezometers


\.t2._f au-'t __
Figure 15. Drainage gallery and connection with asphalt face (preliminary design of
Montanejos dam).

t--- 6.0_0__________~~~

4 1
I +- 2. 8 6'----~"-i
020/0.25

.. I~-t 90
iO.
-.r--

Figure 16. Slab-type plinth at Huesna asphaltic face dam (Spain).


636

Filling of mastic coot (1.5 kg/m2 )

o .,. 0 "" 0

Connection with,
water-s top 0' o
o. o . : 0' 0" o
0... 0'.' ..
. \l
o,.c o ( ) '
o C u, o 0:
Specia I mastic'
""6 em binder layer (0-16m m)
"'"Scm binder layer(5 -22mm )
"~.~inder layer (IS -32 ~~
" ~ionic eiTlulsion primJ..n~~~_ kg/m2)
Geotex tile HoteJit type 30/13
Copper water - stop 0.2 mm x 2 m

asphal t face
Figure 17. Conne ction of Martin Gonza lo plinth (coffer dam) with
(courtesy of Confederaci6n Hidrografica del Guadalquivir, Spain).

This dam had been previously constructed with a thin plastic membrane.
Chapte r 7,
Owing to the large compressibility of rockfill (9.8 MPa) as indicated in
ons) in the semiper-
to a problem of internal erosion (lack of self-filtering conditi
there was a
vious materi al, and to some in situ permea bility of the membr ane,
rupture in it, and it was decided to replace it by an asphaltic face.
it was four
The membr ane was really a landm ark for a dam of this height, and
before taking the decisio n of
times cheape r than the asphal t face. It was stated,
have produc ed a
employing this membrane, that if it had to be substituted it would
settlem ents.
prestressing effect in the rockfill, and the new facing would suffer less
ent it was decide d to take precau tions in the
Notwithstanding, due to the preced
17. A double line of defenc e against
union with the plinth as indicat ed in figure
imperv ious asphalt ic concrete.
leakage was prepared: a copper water-stop and the
with the plinth. A
The imperv ious layer was strong ly thicke ned in the union
geotextiles was placed inside the binder layer to reinforce it.
637

Figure 18 shows the Hatelit geotextile placed, the paving machine placing the
upper binding layer, and the copper water-stops to the right.

Figure 18. Martin Gonzalo repair. Contact with plinth (courtesy of Strabag).
638

Figure 19 shows placement of impervious layer at contact with slab-type plinth at


abutment.

Figure 19. Martin Gonzalo repair. Placement of impervious layer at slab-type


plinth of right abutment during 2nd phase construction of facing (courtesy of
Strabag).

6. Finite element computations

According to our records the first FE calculation of an asphaltic concrete face dam
was performed by Saura (1979).
The three-dimensional FE method is described by the authors and was applied to
Venemo's Dam (v. table I).
The facing was assumed linear elastic.
Several authors attributed to asphaltic concrete moduli of linear deformation
from 10 to 100 MPa (v. Hasegawa and Kikusawa, 1988; Ishii and Kamijo, 1988) and
Poisson's ratio of 0.45.
According to the tests carried out by Kjaernsli et al. (1966), the modulus of linear
639

deformation is very sensitive to bitumen content, strain rate, temperature and


confinement.
With bitumen of penetration 80/100, a bitumen content of 9%, a temperature of
25C, a strain rate of 1.9xlO-6/s and zero confinement, a modulus of 0.4 MPa was
found. This shows that this combination shoud not be used.
For a bitumen content of 7% (finally adopted in Venemo), the modulus may
range from 10 to 75 MPa as temperature decreases from 25 to 2C and strain rate
increases ten-fold.
For that reason, it is very difficult to establish a value of E.
What was done in the calculation of Venemo's dam was to establish a range of
possible variation of E (7.5 to 75 MPa).
The dam was calculated for water load.
For the rockfill the following parameters were chosen:

E = 30MPa
\! = 0.25

This combination gives Eoed = 36 MPa, only slightly larger than the average
value measured during construction (v. table I).
With the parameters indicated, the agreement between measured and calculated
displacements of the facing is excellent:
Maximum calculated displacement normal to the face 11 cm.
Measured displacement: 12 cm.
Three calculations of the facing have been made according to the parameters of it
indicated in table II.
Table II. Asphaltic concrete parameters

I E I
Case I I \!
I(MPa)1
I I
1 175 10.15
I I
2 I 75 I 0.25
I I
3 I 7.5 I 0.25

The results are shown in figures 20 to 22. The following comments can be made.
1. When Poisson's ratio decreases, tensile stresses increase (compare cases 1 and
2).
2. When E decreases from 75 to 7.5 MPa, tensile stresses nearly disappear (compa-
re cases 2 and 3), except at the upper part of facing where they are quite small
640

(2 kPa maximum).

E"75MPa
-;" 0.25
: TENSION

; COMPRESSION

Figure 20. Principal stresses (kPa) in the plane of the facing for full reservoir.
Venema's Dam. Case 1, table II. Justo and Saura (1981).

t
.6
~8 t '7
~3 ~I +-l-!4 .-1!6 .fo
~
4!~ ~~4
....-
,.-7 ... 0

--r-20
'9 , 6 ~7 ~ 7 .7 .. 10 ~ ~ 13
'" 10 3,;:'2

r
~-3

v
~17 ~12. -~r6 -~3 -~2 ... 7 ~3
.. 32.
742. ~ ~ 41 37 --34 -fz9
~
~ 4
.. 2.6
~
--r21
13 ~
-~9
26

-*
, 65
19~
-'f-67 -~ 63
-3
57 46 .'
~6 /' '.,23 53,~6

'66
B~3
r-76 - ~
.. 62 .'4 21 .- 183";..(7
-+67 t 7 r41 - ..\, 22 4 1-
.. 66
-""-"0 --r-79
.56
" 30
~
25
_~21
"4~
_,.54!
1~~37 ~
..

1--.52
33 --1rs86 15 /"'..
"'"32J 3 6
/Y
rnA 47
~'O - ..... 9
E=75 MPa
145%'7i~
-<
.. 24 I: 8 65

.
-"'-,22 MAXIMUM TENSILE i = 0.15
.30 STRAIN 0.39 % t..
113~6'
~
TENSION
~77 ,

tl'7,~ COMPRESSION
----48
*
I
.. ~z

Figure 21. Principal stresses (kPa) in the plane of the facing for full reservoir.
Venema's Dam. Case 2, table II. Justo and Saura (1981),
641

E=15 MPa
y = 0.25
STRAIN 0.41 % t TENSION

~ COMPRESSION

Figure 22. Principal stresses (kPa) in the plane of the facing for full reservoir.
Venemo's Dam. Case 3, tabla II. Justo and Saura (1981).

This reflects the fact that in cases 1 and 2 the facing is more rigid than the rockfill,
and in case 3 it is more flexible.
The distribution of tensile stresses near the abutments and the crest of the dam
correspond to zones at which the joints of concrete faces usually open (Dix,
CogsweeJ, Salt Springs, Lower Bear No.1, Wishon, Paradela, Ishibuchi, Nozori).
The maximum extension strain always occurs at the same place, and has nearly the
same value (around 0.4%), as it is mainly defined by the compressibility of rockfill,
although in the case of figure 22 could not produce a crack because it is associated
with compressive stresses.
In test carried out at the laboratory of duration 3 months (the reservoir was filled
in 4 months) the cracking strain was 1.3%. This explains the good behaviour of the
dam.
The maximum compressive strain was around 0.14%, and was, so, clearly
allowable (v. Kjaernsli et al., 1966).
The following statements can be made:
1. A three-dimensional calculation is necessary to find the stresses and strains in
the plane of the facing.
2. The contac of the facing with the abutment is the most critical zone as far as ten-
sion stresses are concerned. The slope of the abutment will be, so, very important.
3. The relative stiffness facing-rockfill defines the appearance of tensile stresses at
the contact of the facing with the abutment. When the facing becomes more
642

flexible than the rockfill, tensile stresses nearly disappear.


4. The characteristics of the facing have little influence in its deformations, as they
are mostly defined by the deformability of rockfill.
5. It is neccesary to further deepen in the knowledge of the rheological characteris-
tics of asphaltic concrete.

7. Upstream slope

Most of the asphalt face rockfill dams have upstream slopes ranging from 1: 1.6 to
1:2.0 (80% in table I).
Three dams in table I have slopes around 1:2.25 (7.5%)
As indicated already in the Introduction the former asphalt face rockfill dams had
rather steep slopes, from 0.7: 1 to 1: 1. So as to be able to endure the shear stresses
applied by such steep slopes, the bitumen penetration was rather low (20-30 in
Ghrib and Bou Hanifia; 50-75 in Radoina).
For slopes from 1:1.3 to 1:1.5 penetrations 60-70 have been used in Taum-Sank
and Salagou, and 40-50 in Legadadi.
Most of the penetrations range from 40-60 to 80-100 (v. Icold, 1982).
Icold (1982) indicates the following:
It could be said that the maximum slope limit compatible with the construction of
this type of facing is 1:1.5, which is the limit for the stability of the hot mix on the
inclined plane, before and after compaction, and for a safe foothold for workers
without provision of special devices.
Figures 6, 8, 9, 18 and 19 show the work on a 1:1.5 slope.

8. Performance of asphaltic concrete facing rocktill dams

Table I collects many outstanding asphaltic face rockfill dams, on most of which its
performance is known.
There are only two failures.
The first is Mulungushi Dam, in which probably the main cause is the large
compressibility of the dumped rockfill, and the resulting large post-constructive
settlement (1.40 m). The extremely high leakage (7.1 m3 /s), rupture of the facing
and important repair (v. fig. 1) deserve the denomination of failure.
The second failure is Scotts Peak Dam (fig. 23). A crack, more than 50 mm wide
at some places, appeared mainly at the contact between the rather stiff gravel toe
and the more compressible argillite rockfill.
Cracks were filled with gravel and sand, and butyl rubber patches applied on
them. Finally a gravel blanket with 15% below No. 200 sieve was placed on the
643

lower half of the dam (fig. 24).

Max. Flood Level 3097


f'l.:-~-~
i
--
Zone 2 A
i -100 mm Crushed Rock
I 23H

H ~
I 1-7
zonel3A

117

Figure 23. Scotts Peak Dam (Fitzpatrick, 1976).

Bituminous concrete face

Zone 3A

Figure 24. Repair of Scotts Peak Dam (Fitzpatrick, 1976).

Cracks have also appeared at the union of facing with culverts (Wins car dam) or
galleries (Alesani).
We have also described in the Introduction the sliding of the porous reinforced
concrete slabs and the problem of blisters (Bigge, Shibianyu).
Some problems have arised at the contact with the plinth (Grosse, Licheyu,
Hochwurten).
Notwithstanding that, the general performance is excellent. With the exception of
644

the two indicated failures, and the problem of Winscar, in no case has leakage due
to the facing exceeded 7 lis.
The leakage of EI Siberia (15 I/s) was due to a rupture of the plinth (fig. 9) with
no relationship at all with the facing.
At Ghrib Dam self-healing reduced leakage from 25 to 0.1 1/ s.
Long term behaviour of asphaltic concrete is good (v. Weinhold and Haugh,
1988).
Some dams included in table I have suffered important post-constructive
deformations (Ghrib, Prims, Steinbach), and notwithstanding that the performance
was excellent.
Figures 25, 26 and 27 show the aspect of two asphalt face dams, whose performan-
ce is, up to now, excellent.

Figure 25. Martfn Gonzalo asphalt face dam during impounding (courtesy of
Strabag).
Figure 26. Huesna Dam at the end of construction. Designer: Engineer B. Bayan.

'J>
'"+-
646

Figure 27. Huesna Dam. Aspect of facing after first drawdown (courtesy of Stra-
bag).

9. The future of asphaltic concrete facings

Asphaltic concrete facing is an excellent impervious element for dams, due to its
flexibility, speed of construction, self-healing properties and easiness of repair (v.
Ditter and Haug, 1989).
Notwithstanding its future is much related to the price of oil.
In the 1950's an asphaltic concrete facing was much cheaper than a reinforced
concrete one (v. Scheidenhelm and Snethlage, 1960). In the 1980's the cost of an
asphaltic concrete facing in Spain was double the price of a reinforced concrete
deck.
In 1988, in Italy bituminous materials were in the first position among the various
types of upstream impervious elements.
At the moment of writing this Chapter (September 1990) the Gulf crisis throws a
shade of incertitude about the price of asphaltic concrete.
Asphaltic concrete linings are fully compatible with drinking water (v. SchOnian,
1989).
647

The election between asphaltic concrete and reinforced concrete will depend
mainly upon the importance of losses of water and deform ability of rockfill. When
water losses are important and only low grade rockfill is at hand, asphaltic concrete
will probably be a better election owing to its water-tightness and flexibility.
648

References

1. Adalid, J.L., Alonso Franco, M. Nieto Cufi, R, 1973. "Comentarios so-


bre las presas de materiales sueltos con pantalla en Espana". MOPU,
Madrid.
2. Alonso Franco, M., and Soriano, A, 1988. "Plinths in dams with water-
tight facing. Design, construction and performance". 16th Congo Large
Dams, San Francisco, 2:871-900.
3. Daicho, A, 1988. "Design and monitorning of Tataragi dam". 16th Congo
Large Dams, San Francisco, Vol. 2, 205-226.
4. Delgado, J., 1988 "The asphalt facing of the rockfill section of the Negra-
tin dam and its joint with the concrete section". 16Q Congo Grandes Presas,
San Francisco, Vol. 2, 803-814.
5. Ditter, K and Haug, W., 1989. "Refurbishment of bituminous slope lin-
ning on dams and pumped-storage reservoirs". Asphalt-Wasserbau lIla.
Teil A Strabag, p. 43-48.
6. Fabian, E., and Ditter, K, 1988. "Criteria for judgement of the aging be-
haviour of asphaltic surface linings and their influence on the repair and
regeneration of pumped storage reservoir ". 16th Congo Large Dams,
2:375-397.
7. Fitzpatrick, M.D., 1976. "Scotts Peak Dam. Cracking of bituminous con-
crete facing". 12th Congo Large Dams, Mexico, Vol. 5, 54-60.
8. Groupe de Travail du Comite Fran<;ais des Grands Barrages, 1988. "En-
seignements tires de la construction, an cours de ces vingt dernieres
annees, de barrages a masque en enrobes noirs" 16th Congo Large Dams,
San Francisco, Vol. 2: 77-94.
9. Hasegawa, T. and Kikusawa, M., 1988. "Long-term observation of as-
phaltic concrete facing dam". 16th Congo Large Dams, San Francisco,
Vol. 2, 205-226.
1O.ICOLD, 1982. "Bituminous concrete facings for earth and rockfill dams".
Bulletin 32a.
11. Ishii, K and Kamijo, M., 1988. "Design for asphaltic concrete facing ofSabi-
gawa upper dam". 16th Congo Large Dams, San Francisco, 2:327-357.
12. Justo, J.L., 1968. "Deformaci6n de las Presas de Escollera". Fundaci6n
Juan March, Madrid.
13. Justo, J.L., 1973. "The cracking of earth and rockfill dams". 11th Congo
Large Dams, Madrid, 4:921-945.
14. Justo, J.L. and Saura, J., 1981. "Behaviour of Venemo Dam by three-
dimensional FE". 11th ICSMFE, 3:449-452.
649

15. Kjaernsli, B., Mourn, J. and Torblaa, I., 1966. "Laboratory tests on as-
phaltic concrete for an impervious membrane on the Venemo rock-fill
dam". N.G.I:, 69:17-26.
16. Koenig, H.W. and Idel, 1967. "Deformation and loading of a rockfill dam
with bituminous surface membrane". 9th Congo Large Dams, 3:701-712.
17. Legge, G.H.H., 1970. "Mulungushi and Mila Hills dams operation and
maintenance". 16th Congo Large Dams, Montreal, 3:71-90.
18. Matsumoto, N., Yasuda, N., Ogawa, M. and Iwata, M., 1985. "Investiga-
tions of cracks in an asphalt concrete facing and comparison between
observed cracking and predicted behaviour by earthquake analysis". 15th
Congo Large Dams, Lausanne, Vol. 1:559-578.
19. Penman, AD.M. and Charles, J.A, 1985. "Behaviour of rockfill dam with
asphaltic membrane". 11th Congo ICSMFE, San Francisco, 4:2011-2014.
20. Routh, C.D., 1988. "The investigation, identification and repair of the
asphaltic concrete facing of Winscar dam". 16th Congo Large Dams, 2:
655-677.
21. Saura, J., 1979. "Estudio tridimensional de tensiones y deformaciones en
pres as de materiales sueltos". Ph. D. Thesis. Poly technical University of
Madrid.
22. Scheidenhelm, F.W., Snethlage, J.B. and Vanderlip, AN., 1960. "Mont-
gomery dam with asphaltic concrete deck". Trans. ASCE, 125:2B:431-464.
23. Schewe, L. and Geiseler, D., 1985. "Application of asphalt concrete in
sealing of embankment dams". Geotechnik, Special Issue 1985,68-72.
24. SchOnian, E., 1989. "Waterproofing of dams and reservoirs by asphaltic
concrete, its practical performance and compatibility with drinking
water". Asphalt-Wasserbau 11/1. Teil A Strabag, p. 25-36.
25. Sherard, J.L., Woodward, R.J., Gizienski, S.F. and Clevenger, W.A,
1963. "Earth and Earth-Rock Dams". Wiley, N.Y.
26. Smith, L.D., Macpherson, H.H. and Oechsel, R.G., 1988. "Long term
performance. Asphalt lineal upper reservoir Seneca pumped storage
station". 16th Congo Large Dams, San Francisco, 2:501-523.
27. Steffen, H., 1976. "The experience with impervious asphaltic elements
and the conclusions for their design". 12th Congo Large Dams, Mexico,
1:395-405.
28. Strabag. "Asphaltic concrete sealings of hydraulic structures".
29. Tschernutter, P., 1988. "Experience gained with asphaltic concrete fa-
cings on high-level emb ankment dams of the Fragant group of power
schemes". 16th Congo Large Dams, San Francisco, 2: 1105-113l.
30. Van Asbeck, W.F., 1964. "Bitumen in Hydraulic Engineering". Vol. 2
Elsevier, Amsterdam.
650

31. Weinhold, R. and Haug, W., 1988. "The influence of large deformations
on asphaltic cores and membranes -examples of the rockfill dams of
Breitenbach and Prims-reservoir, FRG". 16th Congo Large Dams, 2:399-
414.
32. Yang, Q, Sun, Z. and Ding, P., 1988. "Asphalt concrete facing for rockfill
dams built by directional blasting". 16th Congo Large Dams, San Francis-
co, Vol. 2,1091-1103.
CLOSING SESSION
F. FREDERICO

I appreciate very much the opportunity of making some brief


final considerations.
On behalf of the participants in this Course, I wish firstly
to express our deep gratitude to all who made such an
important event possible, by organizing this Course devoted to
a better understanding of the numerous, and often fundamental,
aspects of the problems related to the design, construction
and monitoring of rockfill structures. In this respect, there
are many people deserving our thanks: I would can just mention
Prof.Oliveira and Prof. Veiga pinto, besides Prof. Maranha das
Neves, among them.
Before making some considerations on the course, let me
repeat the opinion clearly exposed by Prof. Naylor yesterday
night, which I should try to summarize by quoting Prof.
Brauns'words after the song "Coimbra", that ended the banquet
at Palace Hotel, in curia: "I will never forget" (very simple
but effective words). These words hold not only for the
marvellous entertainment by players and singers, including
Professors, but even for each of the components of the whole
course i.e., lectures and lecturers, general organization,
local assistance, technical visits, tourist trips, dinners,
and so on: everything had been planned and developed in the
best imaginable way.
The importance of the matter is evident from the exhaustive
lectures and from the contributions presented by the speakers.
The most advanced methods of analysis for the design of
rockfill structures were exposed; experimental results were
submitted and clearly explained; interesting and well
documented cases of rockfill dams were presented and analyzed;
the role of construction procedures and the role of monitoring
were highly and rightly emphasized.
The interest on the numerous subjects is easily recognizable
through the deep and fruitful discussions that took place
after each lecture.
Lectures and discussions have often widened the field of the
current knowledge on this fascinating subject and suggested
new
ideas for further research work. This is, in my opinion, a
very important point: not only new theoretical and
experimental results, but also new lines of research were
proposed.
I can just mention some few examples, but many others might
651
652

be easily recalled.
The first one regards the concept of limit state design and
the use of partial safety factors. In his lecture, Prof.
Maranha das Neves pointed out, among many different arguments,
such as the different phases of dam life, the role of filters,
the importance of parametric study of the dam behaviour, that
analytical procedures for practical and current applications
to limit state of rockfill structures are lacking.
The second example concerns the phenomenon of collapse of
rockfill, which was analyzed from different points of view by
Professors Justo, Charles, Parkin, Veiga Pinto.
For both examples the speakers recalled that further
research efforts are still needed. To this purpose, it is
worth mentioning, as Prof. Veiga Pinto and Prof. Maranha das
Neves did in their panel discussion on "Rockfill Future", that
the mechanical behaviour of rockf i 11 may also be analyzed
taking into account the peculiarity of this material: i.e. its
granular nature. According to this line of thought, the
analysis should be carried out based on the principles of the
Mechanics of Particulate Materials, taking in due account the
main geometrical (average diameter, shape) and mechanical
(modulus of elasticity and Poisson's ratio) properties of the
single grains as well as the fabric of the medium.
At the end of this very important and interesting course we
are convinced that the existing knowledge is more firmly
founded on Soil Mechanics principles and new paths have been
discovered; as a consequence, new lines of researchs can
positively be followed.
Many thanks again.

J. ANDREW CHARLES

Speaking on behalf of the lecturers I should like first of


all to congratulate LNEC on the selection of this topic
"Advances in Rockfill Structures". In the past rockfill has
received less attention from geotechnical engineers than its
widespread use and economic importance warrant. I think that
this is largely because it is a large, crude material that
does not have the decency to conform to fashionable soil
models. (AS Dr Maranha das Neves reminded us at the beginning
of the ASI, the particle size distribution of rockfill can
change significantly during the course of a laboratory test).
The long standing neglect of rockfill has been remedied by
this NATO ASI.
Secondly I should like to congratulate the participants for
participating: not just being present, but actively
discussing, making presentations and, where necessary,
disagreeing with lecturers. This has given the ASI a depth
which would otherwise have been lacking, however informative
the lectures might be.
I believe that for all of us it has been a profitable two
weeks. We have seen photographs of many rockfill dams
including earth core, concrete faced, asphaltic faced and
asphaltic core dams: very memorably rockfill dams in Brazil,
but also Portugal, Mexico, UK, USA, etc. We have heard a
little about other types of rockfill structures: roads,
breakwaters, opencast mining backfills. We have learned of
testing techniques in Germany, Portugal, UK. We have been
informed about analytical methods. We have seen photographs of
liquefaction failures - but not in rockfill! Hydraulic as well
as geotechnical properties of rockfill have been reviewed.
Last weekend on the field visit we could actually touch real
rockfill!
In the session on "Rockfill future" Professor Nelson Pinto
reminded us that the past is easy to analyse, the future is
impossible to predict and that economists continually
demonstrate these truths. Nevertheless certain technical
issues are clearly going to need further work in the future.

Collapse compression: one of the most important facets


of rockfill behaviour, but the least well understood.

The use of index properties of the parent rock material


to predict fill behaviour: more work is needed to
realise this highly desirable objective.
The importance of details in rockfill structures: e.g.
problems with concrete faced rockfill dams occur at the
perimetric joint.
A growing interest in seismic behaviour.
The use of low grade rockfills: when is rockfill no
longer rockfill?
An incidental benefit of the ASI has been the opportunity to
learn more about the excellent work being carried on in the
Geotechnique Department of LNEC and to meet many members of
the staff. Not only the Geotechnique Department, but also the
shaking tables in the Structures Department would be of
interest to many. We have had the opportunity to meet and walk
with Dr Oliveira, Deputy Director of LNEC. For those of us
from European countries this has been particularly helpful as
increased collaboration among European countries in the next
few years seems almost certain to occur.
Finally on behalf of the lecturers I shooul like to thank
the staff of LNEC who have looked after us so well. We are
conscious of many who have helped us, but one cannot mention
everyone by name. We must thank by name the organisers of the
654

ASI Dr Maranha das Neves and Dr Veiga Pinto. Our grateful


thanks to you all for your help and for your unfailing
personal kindness and courtesy that have helped to make this
such a memorable occasion.

E. MARANHA das NEVES

I will not bore you with a long speech but it is with great
pleasure and gratification that I am addressing some words to
you.
Concerning the Course itself I think that the summing-up
was so well done by Dr. Andrew Charles and Dr. Francesco
Federico that the only thing to be said is to thank their
kind and generous appreciations.
Nevertheless I must also thank to all those who assisted us
in organizing this Course:

- NATO through Dr. Veiga da Cunha, Head of the Scientific


Affairs Division

- LNEC through Prof. Ricardo Oliveira for his support,


without which the ASI would not be what it really was~
and also the staff of LNEC for their dedication and
profissional competence
The Lecturers for their good will and efforts in
preparing and presenting their lectures.
Allow also me to make special mention of our Lecturer Veiga
Pinto. I ask for him an applause for his diligence and great
enthusiasm from the first minute.
As all of you are aware, LNEC is a very big research
institution in civil engineering, with a long tradition but
always trying to evolve to different and less known areas.
As an example, LNEC, together with other research
institutions and Universities is supporting the foundation of
a research center where unsuspected plastic and aesthetic
aspects in many domains of science and technology - why not
also rockfills? - will be exploited and investigated.
A very small and modest example of this type of experience
is included in a record that will be given to you at the end
of this closing session. A question, (one among many others)
can be raised: Does the deceitful non-sense of the sound
sequences that you will hear have any conexions with the
strange and unpredictable creep movements in rockfill? And
may the apparent chaos that prevails in a set of rockfill
particles have some undiscovered order behind it? Is there
something beyond harmony? Are there laws beyond order?
655

A very important subject will be the output of the Course,


i.e. the publication of the information given to us during the
last two weeks. We assure you that we will do our best to make
that information availade to you very soon.
Finally I must say that I would be lying if I told you that
we have not worked very hard to accomplish this Course. But on
behalf of all that worked with us I can assure you that we
will feel highly rewarded if, when returning to your
institutions, your universities, yours firms, in Portugal or
abroad, you will keep good memories of the days spent in this
Course on Advances in Rockfill structures.
LIST OF PARTICIPANTS

Lecturers

A. Parkin Monash University


senior Lecturer Clayton, Melbourne
Victoria, 3168
AUSTRALIA
A. Veiga Pinto Lab. Nac. Eng. civil
senior Research Officer Av. do Brasil, 101
1799 LISBOA CODEX
PORTUGAL

D. Naylor University College of


Senior Lecturer Swansea
Depart. of civil Engine-
ering
Singleton Park,
Swansea, SA2 8PP
U.K.

E. Maranha das Neves Lab. Nac. Eng. civil


Head Geotechnical Department Av. do Brasil, 101
1799 LISBOA CODEX
PORTUGAL

E. Yanagisawa Tohoku University


Professor Depart. of civil Engine-
ering
Faculty of Engineering
Aoba, Sendai 980
JAPAN

H. Evrard Laboratoire Regional des


Head of the Rock Mechanics Ponts et Chaussee
Group CETE de Lyon
109, Avenue Salvador-
Allende CSE Ng 1 - 69674
BRON CEDEX
FRANCE
657
658

J. Brauns Karlsruhe University


Head of section of Soil Postfach 6980
and Rock Mechanics D - 7500 Karlsruhe
GERMANY
J. Charles Building Research
Geotechnics Division Establishment
Garston Watford WD2 7JR
U.K.
J. Delgado Rodrigues Lab. Nac. Eng. civil
principal Research Officer Av. do Brasil. 101
1799 LISBOA CODEX
PORTUGAL
J. Justo Alpanes E. T. S. Arquitecture de
Senior Lecturer Seville
Av Reina Mercedes sin
41012 Seville
SPAIN
N. Sousa Pinto Av. Vicente Machado, 2340
Consulting Engineering 80430 Curitiba - PR
BRASIL

R. Martins Lab. Nac. Eng. civil


principal Research Officer Av. do Brasil, 101
1799 LISBOA CODEX
PORTUGAL

Student Participants

A. B. Moffat University of Newcastle


Senior Lecturer Upon
Tyne - Department of civil
Engineering
Drummond Building
Newcastle Upon Tyne NE1 7RU
U.K.

Ahmet Oguz Tan Technical University of


Associate Professor Istambul
I.T.U. Insaat Fakultesi
Maslak - 80626 Istanbul
TURKEY
Ana A. M. Quintela Cruz Lab. Nac. Eng. civil
Research Assistant Av. do Brasil, 101
1799 LISBOA CODEX
PORTUGAL
659

Armindo Ferreira Lab. de Eng. civil de


Director of Lab. de Eng. Cabo Verde
civil de Cabo Verde CEP 114 - Cidade da Praia
CABO VERDE

Armindo Gomes da Silva Lab. Eng. civil Angola


Research Engineer caixa Postal 1627 -
Luanda
R. P. ANGOLA

Aysen <;:elebi Technical University of


Assistant Istambul
Faculty of Civil Engine-
ering
Ayazaga Istanbul
TURKEY

Bilge Siyahi Anadolu University


Assistant Muh. Mim. Fak. Ins. Bol.
Eskisehir 26010
TURKEY

Carlos dos Santos Pereira Instituto Superior


Assistant Tecnico
Av. Rovisco Pais
1000 LISBOA
PORTUGAL

Carlos Quadros Eduardo Mondlane


Director of Center for University
Eng. Studies Faculty of Engineering
P.O. Box 257 - Maputo
MO<;:AMBIQUE

Da-Mang Lee Cambridge University


Research Student Department of Engineering
Trumpington Street
Cambridge CC2 IPZ
U.K.

Decio Mattar Junior IPT


Geotechnical Engineer cidade universitaria
CEP 05508 - Sao Paulo, SP
CP 7141 - CEP
01051 BRASIL

Domingos Torres Guimaraes Lab. Eng. civil Angola


Research Engineer Caixa Postal 1627 -Luanda
R. P. ANGOLA
660

Efrossini C. Kalkani National Technical Asso-


Professor ciate University
civil Engineering,
Sevastoupaleos, 50
11526 Athens
GREECE
Ernesto Oliveira Domingues BRISA
Supervisor Engineer Fiscaliza9ao de T. Novas/
/Fatima/Leiria
(DTF-L) - Cardosos
- Arrabal
2400 LEIRIA
PORTUGAL

Francesco Federico University of Rome "Tor


Researcher in Geotechnics Vergata"
Department of civil
Engineering
Via E. Carnevale, 00173
Rome
ITALY

Hakan Sirin Dokuz Eylul University


Research Assistent Faculty of Engineering
and Architecture - Depart.
of civil Engineering
35100 Bornova/Izmir
TURKEY

Huseyin Yildirim Technical University of


Associate Professor Istambul
civil Engineering Fakulty
Ayazaga Istanbul
TURKEY

Ian Christopher pyrah Sheffield University


Lecturer Department of civil and
Structural Engineering
Mappin Street
Sheffield Sl 3JD
U.K.

Jesus Yague Cordova Min. de Obras Publicas e


Safety Dam Engineer Urbanismo - Direcciom
General Obras Hidraulicas
Area Vigilancia de Presas
Paseo Castellana, 67
SPAIN
661

Joao Luis Andrade Cavilhas HP - Hidrotecnica


Geotechnical Consultant Portuguesa
Rua da Gui~e, Prior Velho
2685 SACAVEM
PORTUGAL
Joao Manuel Barros Gomes D.G.R.N.
Supervisor Engineer Av. Almirante Gago
Coutinho, 30
1000 LISBOA
PORTUGAL

Joao Mateus da Silva Lab. Nac. Eng. civil


Research Assistant Av. do Brasil, 101
1799 LISBOA CODEX
PORTUGAL

Jose Antonio Mateus de Brito CENOR


Geotechnical Consultant Av. Almirante Gago
coutinho, 133
1700 LISBOA
PORTUGAL
Jose Couto Marques Fac. Eng. Universidade do
Associate Professor Porto - FEUP
Rua dos Bragas
4099 PORTO CODEX
PORTUGAL

Jose Luis Ribeiro dos Santos JAE - Direc9ao de Estradas


Deputy Director of Santarem de Santarem
Road Office 2000 SANTAREM
PORTUGAL

Juan Carlos Azanedo CEDEX - Lab. de Geotecnia


Geotechnical Engineer CI Alfonso III SIN
Madrid - 28014
SPAIN

K. Kast University Karlsrhue


Chief Researvh Engineer Postfach 6980
D - 7500 Karlsruhe 1
GERMANY

Lucia Maria Fonseca de Almeida Hidroprojecto


Geotechnical Engineer Av. Marechal Craveiro
Lopes, 6
1700 LISBOA
PORTUGAL
662

Luis E. Montanez cartaxo Federal Comission of


Head of the Soil Mech. and Electricity
Found. Office, CFE Augusto Rodin, 265
0~720 Mexico D.F.
MEXICO
Luis Joaquim Lemos Sec~ao Aut6noma de En-
Associate Professor genharia
Fac. de Ciencias e
Tecnologia
Universidade de Coimbra
Largo Marques de Pombal
3049 COIMBRA CODEX
PORTUGAL
Luis Miguel de Cruz Simoes Faculdade de Engenharia
Associate Professor Universidade de Coimbra
Largo Marques de Pombal
3049 COIMBRA CODEX
PORTUGAL

Luis R. Morales Virgen ALSYR, S.A.


Consultant Civil Eng. TEPIC 39, Colonia Roma
Geotechnique M~xico D.F.- C.P. 06760
MEXICO

Malte Cederstrom Vattenfall


Dam Safety Engineer Swedish State Power Board
S - 16287 Stockholm
SWEDEN

Manuel Romana Ruiz universidad Pollitecnica


Professor of Geotechnical de Valencia
Engineering 46071 VALENCIA
SPAIN

Maria Emilia Borralho D.G.R.N.


Geotechnical Engineer Av. Almirante Gago
Coutinho, 30
1000 LISBOA
PORTUGAL

Maria Eugenia Monteiro da Rocha EDP


safety dam Engineer Direc~ao da produ~ao
Hidrau1ica
Largo Dr. Tito Fontes,
n 15
Q

4000 PORTO
PORTUGAL
663

Mario de Quinta Ferreira Departamento de Geotecnia


Assistant Universidade de Coimbra
Largo Marques de Pombal
3049 COIMBRA CODEX
PORTUGAL
Michael D. Pachakis Central Public Works
Research Engineer Laboratory
166, Pineos street
11854 Athens
GREECE
Mittiades K. Zacas PANGEA - Consulting
Director Engineers
131, Kifisias Aven.
11524 Athens
GREECE

Mozart Barbosa Filho universidade Catolica de


Assistant Professor Petropolis
Rua Major Sergio, 140 -
- Mosela
Petropolis RJ - CEP 25675
BRASIL

Niels Johansson Alvkarleby Laboratory


Head of Hydro Power & River Swedish state Power
Hidraulics Board
S - 81071 Alvkarleby
SWEDEN

Nooy Van Der Kolff Delft Geotechnics


Project Engineer P.O. BOX 69
2600 AB Delft
THE NETHERLANDS

octavio Machado Filho Geotecnica S.A.


Geotechnical Engineer Rua Moura Brasil, N 44
Larangeiras CEP 22231
Rio Janeiro - RJ
BRASIL

Ricardo Oliveira Lab. Nac. Eng. Civil


Deputy Director Av. do Brasil, 101
1799 LISBOA CODEX
PORTUGAL
664

Richard M. Doake Mott Macdonald Interna-


Associate Professor tional Lt.
Demeter House
station Road
Cambridge CB1 2RS
U.K.

Serafim Luis BRISA


Head of section of Quality Av. Fontes Pereira de
Control Melo I 6 - 3 0
1000 LISBOA
PORTUGAL

victor Figueiredo de Jesus HP - Hidrotecnica


Geotechnical Consultant Portuguesa
Rua da Gui~e, Prior Velho
2685 SACAVEM
PORTUGAL

W. Sadgorski Water Department of


Soil Mechanics Engineer Baviera
Rotbuchen str. 73
8000 Munchen 90
GERMANY

You might also like