You are on page 1of 287

McGill University

GUIDELINES FOR PRELIMINARY DESIGN OF UNLINED PRESSURE


TUNNELS

By
André J. Rancourt

Supervisor : Prof. Hani S. Mitri

A thesis submitted to the Faculty of Graduate Studies and Research in partial


fulfillment of the requirements for the degree of Doctor of Philosophy

© André J. Rancourt, 2010


Library and Archives Bibliothèque et
Canada Archives Canada
Published Heritage Direction du
Branch Patrimoine de l'édition

395 Wellington Street 395, rue Wellington


Ottawa ON K1A 0N4 Ottawa ON K1A 0N4
Canada Canada
Your file Votre référence
ISBN: 978-0-494-72700-3

Our file Notre référence


ISBN: 978-0-494-72700-3

NOTICE: AVIS:
The author has granted a non- L'auteur a accordé une licence non exclusive
exclusive license allowing Library and permettant à la Bibliothèque et Archives
Archives Canada to reproduce, Canada de reproduire, publier, archiver,
publish, archive, preserve, conserve, sauvegarder, conserver, transmettre au public
communicate to the public by par télécommunication ou par l'Internet, prêter,
telecommunication or on the Internet, distribuer et vendre des thèses partout dans le
loan, distrbute and sell theses monde, à des fins commerciales ou autres, sur
worldwide, for commercial or non- support microforme, papier, électronique et/ou
commercial purposes, in microform, autres formats.
paper, electronic and/or any other
formats.

The author retains copyright L'auteur conserve la propriété du droit d'auteur


ownership and moral rights in this et des droits moraux qui protege cette thèse. Ni
thesis. Neither the thesis nor la thèse ni des extraits substantiels de celle-ci
substantial extracts from it may be ne doivent être imprimés ou autrement
printed or otherwise reproduced reproduits sans son autorisation.
without the author's permission.

In compliance with the Canadian Conformément à la loi canadienne sur la


Privacy Act some supporting forms protection de la vie privée, quelques
may have been removed from this formulaires secondaires ont été enlevés de
thesis. cette thèse.

While these forms may be included Bien que ces formulaires aient inclus dans
in the document page count, their la pagination, il n'y aura aucun contenu
removal does not represent any loss manquant.
of content from the thesis.
The material world is a possible movement
of consciousness and I am choosing moments
out of these movements to bring my actual
experience into manifestation

A. Goswani
ABSTRACT

Unlined pressure tunnels are used to convey water to hydroelectric powerhouses. Usually
each project has a forebay reservoir, an intake, a pressurized tunnel, a surface or
underground powerhouse and a tailrace tunnel or channel. This thesis deals with the
design of unlined pressure tunnels. These were first driven by the Norwegians in hard
granitic rock masses, which would constitute the ideal material for this type of structure.
However the increase in demand for electrical energy has pushed the industry to
implement projects in sites with less favourable geological conditions all over the world.
For each pressure tunnel, there are normally three different types of sections along its
length: the unlined section, the semi-lined section and the lined section. The unlined
pressure tunnel is pervious to a certain extent even if the rock is hard and non-porous and
the rock mass is of good quality. Unlined tunnels are thus prone to hydraulic jacking.
Semi-lined tunnels use unreinforced shotcrete and/ or concrete whereas fully lined
tunnels are defined as those with a steel liner cast in concrete and completed with some
annulus grouting to fill voids in the steel-concrete-rock mass system around the tunnel.
This type of lining is totally impervious but is very expensive to build. Thus, unlined
pressure tunnels, being the most economical, play an important role in the economics of a
hydroelectric power project.

In this thesis, an extensive review of literature is conducted and most published unlined
pressure tunnel design guidelines against hydraulic jacking are reviewed. The thesis
places emphasis on the preliminary phase when no field measurements are available. The
spatial distribution of the minimum stress required to control hydraulic jacking around
unlined pressure tunnel is studied. The topographic effects, the rock cover-to-tunnel
diameter ratio, the presence of geological feature are all investigated using FLAC 2D
code. Minimum stress results are compared to the minimum stress given by the well-
known Norwegian design criterion, and a correction factor called, Cover Alteration Ratio
(CAR) is introduced. Cases where the Norwegian criterion is not adequate are identified
and it is proposed to deal with these situations by adopting a larger factor of safety.
The FLAC 2D analyses show that the presence of topographic features has a great
influence on the minimum stress distribution. Also, that Norwegian criterion is not
adequate when the tunnel diameter is large compared to the rock cover. Finally, the
presence of structural features near the tunnel, greatly change the minimum stress
distribution and thus renders the Norwegian criterion inadequate in those cases.

Based on these results, a design methodology is proposed for use at the preliminary stage
in the form of a factor of safety against hydraulic jacking to be used in the Norwegian
criterion. Limitations on the use of the criterion are presented. The thesis presents
recommendations to orient investigations and for final tunnel and liner design.

ii
RÉSUMÉ

Les conduites en charge sont largement utilisées comme galerie d'amenée pour les
centrales hydroélectriques. D'une façon générale, chaque projet est composé d'un
réservoir, d’une prise d'eau, d’une ou plusieurs conduites forcées, d’une centrale de
surface ou souterraine et d’un canal ou d’une galerie de fuite. Historiquement, les
conduites en charge non revêtues ont été utilisées par les Norvégiens dans des massifs
rocheux de type granitique de bonne à excellente qualité. Cependant, la demande sans
cesse croissante pour l'énergie électrique à travers le monde, a conduit l'industrie à
développer des sites présentant des conditions géologiques de moins en moins favorables.

Trois types de revêtement sont possibles, non revêtus, semi-revêtus et blindés. Même si le
massif rocheux est de bonne qualité, insoluble et peu poreux, les sections de galerie non
revêtues ne sont jamais totalement imperméables. La consolidation des sections non
revêtues est constituée de boulons et de béton projeté appliqué localement selon les
conditions géologiques. Dans le cas où le roc est légèrement soluble, ou poreux, et en
présence d’un massif rocheux de mauvaise qualité, il est requis de couvrir les parois de
béton projeté, ou même de béton coffré. Dans les pires conditions, le béton peut être armé
et constitue ainsi un semi-revêtement pouvant résister aux fortes pressions internes. Ce
revêtement n’est pas complètement étanche puisque les pressions développent des
fissures longitudinales dans le béton qui laisse une certaine quantité d’eau s’échapper
dépendamment de l’ouverture des fissures. Cette ouverture est cependant limitée par
l’armature et la réaction du massif rocheux.

Les galeries blindées sont composées d’un blindage d’acier encastré dans du béton.
L’ensemble est par la suite injecté avec du coulis de ciment pour combler notamment les
vides laissés lors de la mise en place du béton et par le retrait lors de la cure. Ce
revêtement est complètement imperméable et doit être conçu en tenant compte de la
réaction du massif rocheux et également des pressions externes lors d’une vidange. Étant
donné sont coût très élevé, la longueur de cette section blindée est un élément crucial lors
de la conception de la galerie. Cette thèse présente des résultats d’études et des
recommandations concernant le type, et la localisation des divers types de revêtements.

La conception des conduites en charge peut être divisée en trois phases distinctes, la
conception préliminaire, la phase d’investigation et la conception finale. Une ample revue
de la littérature couvrant la plupart des publications concernant la conception des
conduites non revêtues a été réalisée. Cette thèse résume les méthodes et critères de
conception existants, avec une emphase sur l’étape de conception préliminaire alors
qu’aucune donnée d’investigation n’est disponible.

Pour ce faire, la distribution spatiale de la contrainte minimale autour de la galerie aux


environs de la limite du blindage a été étudiée. Le logiciel FLAC 2D a été utilisé pour
modéliser l’effet de la topographie, l’influence du ratio couvert rocheux sur diamètre de
la galerie et la présence de structures géologiques près de la galerie. La contrainte
minimale a été ensuite comparée à la contrainte donnée par le critère Norvégien. Dans
certain cas, les modèles ont montrés que le critère surestimait la contrainte minimale et il
est recommandé, dans ces cas, d’augmenter le coefficient de sécurité.

Les analyses réalisées avec le logiciel FLAC2D ont démontré que la topographie avait
une grande influence sur la distribution de la contrainte minimale. Les analyses ont
également montré que le critère Norvégien n’était pas adéquat quand le diamètre de la
galerie est grand comparé au couvert rocheux. Finalement, la présence de structures
géologiques près de la galerie, altère substantiellement le champ de contrainte ce qui rend
le critère Norvégien inapplicable dans ces situations.

Sur la base de l’ensemble de ces résultats, une méthodologie est proposée pour la
conception préliminaire. Il est proposé d’ajuster le coefficient de sécurité figurant dans le
critère Norvégien pour tenir compte des différents cas problématiques. Également,
certaines limites d’utilisation du critère sont proposées. Finalement, cette thèse présente
des recommandations permettant d’orienter les investigations concernant la conception
finale de la galerie et des sections blindées.

ii
DEDICATION
This work is dedicated to my parents,
Jacques Rancourt and Madeleine Fortin,
with all the affection of a thankful son;
and to my wife Nancy Bertrand and children Camille and Guillaume
with all the love of a husband and father.
AKNOWLEDGEMENTS

I wish to express my sincere appreciation to my supervisor Prof. Hani S. Mitri for


his guidance, direction and availability throughout the work performed during this
thesis. The numerous discussions have helped me refine the approach towards this
engineering geology problem.

I wish to express my gratitude to Hydro-Quebec for their support and permission


to use and publish technical information. I am also extremely thankful to Dr. Ray
Benson, Prof. Einar Broch, Dr. Wynfrith Riemer, Dr. Thomas Doe and Dr.
Alberto Marulanda for their review of the thesis and their precious comments that
have greatly enhanced this thesis.

Finally, I am grateful for the love, understanding and constant encouragement


provided by my wife, Nancy Bertrand, my daughter Camille and son Guillaume.

André J. Rancourt
TABLE OF CONTENTS
Page

ABSTRACT
RÉSUMÉ
DEDICATION
ACKNOWLEDGEMENTS

CHAPTER 1 _____________________________________________________ 1
INTRODUCTION ____________________________________________________ 1
1.1 Background ___________________________________________________ 1
1.2 Study problem _________________________________________________ 5
1.3 Research objective _____________________________________________ 7
1.4 Thesis structure________________________________________________ 8
CHAPTER 2 ____________________________________________________ 10
METHODS OF UNLINED PRESSURE TUNNEL DESIGN ________________ 10
2.1 Types of unlined pressure tunnel failure __________________________ 10
2.1.1 Excessive leakage ______________________________________________ 10
2.1.1.1 High natural permeability ________________________________________ 11
2.1.1.2 Hydraulic failure _______________________________________________ 14
2.1.2 Geomechanical instabilities _______________________________________ 16
2.1.2.1 Tunnel collapse ______________________________________________ 16
2.1.2.2 Surface instability ____________________________________________ 17
2.2 Existing design criteria _________________________________________ 19
2.3 Design methods _______________________________________________ 24
CHAPTER 3 ____________________________________________________ 28
PARAMETERS CONTROLLING UNLINED PRESSURE TUNNEL
BEHAVIOUR_______________________________________________________ 28
3.1 Foreword ____________________________________________________ 28
3.2 Rock mass quality _____________________________________________ 28
3.2.1 Preliminary geomechanical assessment ______________________________ 32
3.2.2 Rock mass investigation _________________________________________ 33
3.3 Hydrogeology ________________________________________________ 34
3.3.1 Preliminary pore pressure estimation ________________________________ 36
3.3.2 Pore pressure measurement _______________________________________ 37
3.4 In-situ stresses ________________________________________________ 38
3.4.1 Preliminary stress estimation ______________________________________ 39
3.4.2 Stress measurements ____________________________________________ 43
3.4.2.1 Stress relief methods __________________________________________ 43
3.4.2.2 Water injection methods _______________________________________ 44
3.5 Rock mass permeability ________________________________________ 50
3.5.1 Open joints and faults ___________________________________________ 52

i
3.5.2 Equivalent fractured continuum ____________________________________ 52
3.5.3 Preliminary permeability estimation ________________________________ 52
3.5.4 Permeability measurements _______________________________________ 53
3.5.4.1 Equivalent continuum _________________________________________ 53
3.6 Slope stability ________________________________________________ 56
3.6.1 Preliminary structural assessment __________________________________ 59
3.6.2 Detailed structural investigation ___________________________________ 59
3.7 Basic parameter interactions ____________________________________ 59
CHAPTER 4 ____________________________________________________ 62
NUMERICAL MODELLING OF SHALLOW UNLINED PRESSURE TUNNEL
___________________________________________________________________ 62
4.1 Numerical modelling method ___________________________________ 63
4.2 Selection of modelling tool ______________________________________ 65
4.2.1 FLAC 2D _____________________________________________________ 66
4.3 Characterization of input data for pressure tunnel models ___________ 66
4.3.1 Model geometry ________________________________________________ 67
4.3.1.1 Tunnel geometry _____________________________________________ 67
4.3.1.2 Tunnel diameter versus rock cover _______________________________ 68
4.3.1.3 Topography _________________________________________________ 69
4.3.2 Structural features ______________________________________________ 72
4.3.3 Structural anisotropy ____________________________________________ 75
4.3.4 Comparison between models with or without structural features __________ 77
4.3.5 Mohr-Coulomb plastic models ____________________________________ 78
4.3.6 Hydro-Mechanical interaction models _______________________________ 80
4.3.7 Boundary conditions ____________________________________________ 81
4.3.8 Rock mass properties ____________________________________________ 83
4.3.8.1 Deformation parameters _______________________________________ 84
4.3.8.2 Hydraulic parameters _________________________________________ 84
4.4 Model parameters _____________________________________________ 85
4.4.1 Isotropic models ________________________________________________ 85
4.4.2 Transversely isotropic models (anisotropic) __________________________ 85
4.4.3 Mohr-Coulomb plastic models ____________________________________ 86
4.4.4 Hydro-mechanical interaction models _______________________________ 86
4.5 Modelling criteria for preliminary tunnel design ___________________ 87
4.5.1 Estimation of the effective rock cover _______________________________ 87
4.5.2 Horizontal stress _______________________________________________ 89
4.5.3 Leakage flow __________________________________________________ 90
CHAPTER 5 ____________________________________________________ 91
MODEL PARAMETRIC STUDIES ____________________________________ 91
5.1 Extent of the effective cover and decompressed areas _______________ 91
5.1.1 Circular versus D-shape tunnel ____________________________________ 91
5.1.2 Tunnel diameter versus rock cover and location of K’ <1 areas ___________ 95
5.1.2.1 Horizontal ground surface ______________________________________ 95
5.1.2.2 Inclined slope _______________________________________________ 96
5.1.2.3 Hill _______________________________________________________ 98
5.1.2.4 Valley ____________________________________________________ 102
5.1.3 Structural features _____________________________________________ 106
5.1.3.1 Distance of a thin soft zone ____________________________________ 106

ii
5.1.3.2 Thickness of the soft zone _____________________________________ 108
5.1.3.3 Variation of elasticity of the structural features ____________________ 110
5.1.3.4 Variation of structural feature orientations (soft zones and hard zones) __ 113
5.1.3.5 Anisotropic models __________________________________________ 117
5.1.3.6 Topographic models with structures _____________________________ 120
5.1.3.7 Plastic models with structures __________________________________ 124
5.2 Rock mass quality effect on expected leakage _____________________ 127
5.3 Conclusions and recommendations from the modelling results _______ 130
CHAPTER 6 ___________________________________________________ 132
PRELIMINARY LAYOUT AND DESIGN GUIDELINES _________________ 132
6.1 Factor of safety assessment ____________________________________ 133
6.2 Integration of numerical results into the preliminary layout _________ 133
6.3 Investigation ________________________________________________ 140
6.4 Final layout _________________________________________________ 143
6.4.1 Erosion lining_________________________________________________ 144
6.4.2 Seal lining and allowable gradient _________________________________ 145
6.4.3 Slope stability ________________________________________________ 146
6.4.4 Other considerations ___________________________________________ 146
CHAPTER 7 ___________________________________________________ 148
CONCLUDING REMARKS_________________________________________ 148
7.1 Numerical modelling study ____________________________________ 148
7.1.1 Spatial variability and parameter estimation _________________________ 148
7.1.2 Limitations of numerical modelling ________________________________ 149
7.2 Design guidelines_____________________________________________ 151
7.3 Summary ___________________________________________________ 152
7.4 Recommendations for future study ______________________________ 155
Thesis contribution and originality ____________________________________ 156

iii
LIST OF TABLES
Page

Table 2.1 – Evolution of cover criteria ________________________________ 20


Table 2.2 - Evolution of design methodologies __________________________ 25
Table 3.1 - Mechanisms of erosion and dissolution ______________________ 29
Table 3.2 - Some available rock mass deformation models from the literature _ 31
Table 3.3 - Textbook density ranges according to rock type _______________ 33
Table 3.4 - Summary of S values from the literature _____________________ 35
Table 3.5 - Structural failure by sliding along slopes ____________________ 57
Table 3.6 - Hydraulic failure along structural features ___________________ 58
Table 3.7 - Summary of coupled permeability models ____________________ 61
Table 4.1 - Summary of numerical modelling studies_____________________ 62
Table 4.2 - Parameters controlling the geometry of the hill/valley models ____ 71
Table 4.3 - Range of rock mass elastic parameters depending
on rock mass quality _____________________________________ 84
Table 4.4 - Isotropic elastic model parameters _________________________ 85
Table 4.5 - Transversely isotropic model parameters ____________________ 85
Table 4.6 - Mohr-Coulomb plastic models parameters ___________________ 86
Table 4.7 - Hydraulic parameters in the models ________________________ 87
Table 5.1 – Results from the hill models (DT = 10 m; C = 120 m) _________ 102
Table 5.2 - Results from the valley models (DT = 10 m; C = 25 m) _________ 106
Table 5.3 - Results from the hill models with a vertical structure
(DT = 10 m; C = 120 m) _________________________________ 122
Table 5.4 - Results from the valley models with a vertical structure
(DT = 10 m; C = 25 m) __________________________________ 124
Table 5.5 - – CAR results for elastic models compared to plastic models ____ 125
Table 5.6 - CAR results for elastic models compared to plastic models _____ 127
Table 6.1 – Propose FS ___________________________________________ 134
Table 6.2 - Preliminary stability verification table______________________ 135
Table 6.3 - Proposed guidelines to evaluate possible effects of natural pore
pressure on the effective driving pressure in tunnels _________ 136
Table 6.4 - List of projects used to calibrate the factors of safety of table 6.1__139
Table 6.5 - Type of investigation recommended for each zone______________141
Table 6.6 - Proposed testing methods related to expected geomechanical
properties_____________________________________________ 143
Table 6.7 – Guidelines for types of erosion lining for estimated rock mass
parameters_____________________________________________________ 144
Table 6.8 - Guidelines for types of seal lining for estimated rock mass parameters
and permeability ________________________________________________ 145
Table 6.9 - Recommended guidelines for allowable gradient and
filling/dewatering rate___________________________________ 147

iv
LIST OF FIGURES
Page
Figure 1.1 – Typical pressure tunnel and shaft schemes
for hydroelectric projects _________________________________ 3
Figure 1.2 – Typical type of pressure tunnel and shaft lining for hydroelectric
projects _______________________________________________ 4
Figure 1.3 – Norwegian criterion (modified from Berg-Christensen
and Dannevig, 1971) ____________________________________ 5
Figure 1.4 – Topographic correction (modified from Broch,1982) ___________ 6
Figure 2.1 – Schematic showing difference between various rock media _____ 14
Figure 2.2 – Slope stability analysis of a rock wedge subjected to water pressure
(modified from Hoek and Bray, 1974) ______________________ 18
Figure 2.3 - Unlined pressure tunnel design methodology (modified
from Merritt, 1999) ____________________________________ 26
Figure 3.1 – Comparison of deformability models _______________________ 32
Figure 3.2 - Typical shallow topographic conditions _____________________ 40
Figure 3.3 - Typical hydraulic testing curve (from Rancourt and Mitri, 2007) _ 47
Figure 3.4 - Typical tendency for permeable and impervious rock masses
(modified from Marulanda et al., 1990) ____________________ 48
Figure 3.5 - Examples of testing situations, a) testing carried from the surface
before tunnel excavation, undrained, b) testing from inside the
tunnel with a drained rock mass __________________________ 50
Figure 3.6 – Typical permeability ranges for rocks and soils (Freeze and Cherry,
1979) _______________________________________________ 53
Figure 3.7 – Typical relations of effective pressure against water flow (modified
from Widmann, 1996) __________________________________ 54
Figure 4.1 – Typical tunnel geometries _______________________________ 67
Figure 4.2 – Tunnel geometries _____________________________________ 68
Figure 4.3 - Models schematic layouts ________________________________ 69
Figure 4.4 – Type of topography modelled _____________________________ 70
Figure 4.5 – Schematic geometry of models with C/DT ratio and topographic
variations ____________________________________________ 71
Figure 4.6 – Schematic geometry of preliminary soft zone models __________ 72
Figure 4.7 – Schematic geometries of preliminary soft zone models _________ 73
Figure 4.8 – Schematic geometries of elasticity models ___________________ 74
Figure 4.9 – Schematic geometry of models with variation of orientation of
structural features _____________________________________ 75
Figure 4.10 – Schematic geometry of anisotropic models (angle ω represent the
angle of E1) __________________________________________ 76
Figure 4.11 – Schematic geometries of hill comparison models ____________ 79
Figure 4.12 – Schematic geometries of valley comparison models with overburden
_______________________________________________________________ 80

v
Figure 4.13 – Schematic geometries of plastic models ___________________ 81
Figure 4.14 – Relation between volumetric strain and permeability ________ 82
Figure 4.15 – Schematic geometry of the hydro-mechanical
interaction models ________________________________________________ 83
Figure 4.16 – Assumed in situ stress pattern to model erosion ______________82
Figure 4.17 – Boundary conditions for hydro-mechanical models ___________83
Figure 4.18 – Minimum stress isocontours and location
of stress alteration from the tunnel_________________________89
Figure 5.1 – Typical FLAC mesh for the circular tunnel __________________ 92
Figure 5.2 – Typical FLAC mesh for the D-shape tunnel__________________ 92
Figure 5.3 – Typical FLAC minimum principal stress contours for a circular
tunnel _______________________________________________ 93
Figure 5.4 – Typical FLAC minimum principal stress contours for a D-shape
tunnel _______________________________________________ 93
Figure 5.5 – C/DT ratio versus the difference in the Cover alteration ratio from
the two tunnel geometries _______________________________ 94
Figure 5.6 – Cover alteration ratio results for horizontal ground models _____ 96
Figure 5.7 – Typical FLAC mesh for the 30° slope models ________________ 97
Figure 5.8 – Typical FLAC minimum principal stress contours for the 30° slope
models ______________________________________________ 97
Figure 5.9 – Cover alteration ratio results for slope models _______________ 98
Figure 5.10 – Typical hill model with correction concept and important
parameters ___________________________________________ 99
Figure 5.11 – Typical FLAC mesh for the 52° hill models _________________ 99
Figure 5.12 – Typical FLAC minimum principal stress contours for the 52° hill
models _____________________________________________ 100
Figure 5.13 – Typical FLAC K’ contours for the 52° hill models (K’ max = 1) 100
Figure 5.14 – Hill models investigated (DT = 10 m; C = 120 m) __________ 101
Figure 5.15 – Typical FLAC mesh for the 52° valley models ______________ 102
Figure 5.16 – Typical FLAC minimum principal stress contours s _________ 103
Figure 5.17 – Typical FLAC K’ contours for the 52° valley models (K’ max = 1)
___________________________________________________ 103
Figure 5.18 – CAR results from the valley models (DT = 10 m; C = 25 m) __ 104
Figure 5.19 – Valley models investigated (DT = 10 m; C = 25 m) _________ 105
Figure 5.20 – Soft zone modelling scenarios __________________________ 107
Figure 5.21 – Typical FLAC minimum principal stress contours for the vertical
soft zone models ______________________________________ 107
Figure 5.22 – Results for the variation of distance of the 3 m thick soft rock zone
(DT = 10 m) _________________________________________ 108
Figure 5.23 - Variation in the thickness of the soft zone _________________ 109
Figure 5.24 - Results for the variation of thickness of the soft zone _________ 109

vi
Figure 5.25 – Layouts of structural feature and different topographies with
variation of elasticity __________________________________ 111
Figure 5.26 – Typical FLAC minimum principal stress contours for E1/E2 = 0.2
models _____________________________________________ 112
Figure 5.27 - Results for the variation of elasticity of the structural feature
according to rock mass host_____________________________ 112
Figure 5.28 – Variation of structure orientations with different topographies 114
Figure 5.29 – Typical FLAC minimum principal stress contours for soft structural
feature models _______________________________________ 115
Figure 5.30 - Results for the variation of orientation of the soft rock 10 m thick
zones _______________________________________________ 115
Figure 5.31 – Typical FLAC minimum principal stress contours for hard
structural feature models _______________________________ 116
Figure 5.32 - Results for the variation of orientation of the hard rock zone __ 117
Figure 5.33 – Variation of the anisotropy orientations with different topographies
(angle ω represent the angle of E’1) ______________________ 118
Figure 5.34 – Typical FLAC minimum principal stress contours for anisotropic
models _____________________________________________ 119
Figure 5.35 - Results for the variation of the anisotropy angle ____________ 120
Figure 5.36 - Results for the variation of the anisotropy intensity ratio (E’1/E’2)
for flat models _______________________________________ 120
Figure 5.37 – Comparison between different hill models_________________ 121
Figure 5.38 – K’ and Z results from the buried valley models (DT = 10 m; C =
25 m) ______________________________________________ 123
Figure 5.39 – Minimum stress results from plastic models compared with
minimum stress from elastic models for soft rock structure ____ 125
Figure 5.40 – Minimum stress results from the plastic models compared with
minimum stress from elastic model for hard rock structure ____ 126
Figure 5.41 – Minimum stress for uncoupled and coupled models _________ 128
Figure 5.42 – Pore pressure for uncoupled and coupled models ___________ 128
Figure 5.43 – Variation of permeability (X 10-12 m2) compared to a constant
uncoupled permeability of 1.0 X 10-12 m2 _________________ 129
Figure 5.44 – Coupling model results _______________________________ 130

vii
LIST OF APPENDICES

Appendix A - Model table parameters


Appendix B – FLAC numerical codes
Appendix C – FLAC models basic geometries and meshes
Appendix D – Minimum stress (σmmin) contour lines showing stress perturbation
and zones where CAR < 1

LIST OF SYMBOLS

αi : dip direction of joint set i [A]


βi : dip angle of joint set I [A]
β: valley wall inclination from the horizontal [degree]
γw : unit weight of water (9.81 kN/m3) [M/T2L3]
γr : unit weight of rock [M/T2L3]
γov : unit weight of overburden [M/T2L3]
δ: angle of the structural feature in numerical models
ε: strain [L/L]
ε0 : initial strain [L/L]
θ: angle of the sliding surface according to horizontal [degree]
λ: joint sets dip angles [degree]
μ : dynamic viscosity of water (10-3 N•sec/m2) [M/TL3]
ν: Poisson ratio
ρw : density of water (1,000 kg/m3) [M/L3]
σ’ : effective stress [M/T2L]
σtot : total stress [M/T2L]
σ’j : effective stress perpendicular to the joint [M/T2L]
σ1,3 : principal stresses [M/T2L]
σ’1,3 : effective principal stresses [M/T2L]
σh : horizontal stresses [M/T2L]
σv : vertical stresses [M/T2L]

viii
σmin : minimum stress [M/T2L]
σmmin : minimum stress obtained from numerical models [M/T2L]
σcmin : minimum stress estimated using Norwegian criterion [M/T2L]
φ: friction angle of the joint [degree]
φ ‘: effective friction angle of the joint [degree]
ω: angle of modulus E1 according to the horizontal for anisotropic models
a: geometrical parameter for valley and hill shape
A: total crack area [L2]
Ac : contact areas inside the crack [L2]
Af : flow cross-sectional area [L2]
B1, B2 : water force ratios; B1 ≈ B2 ≈ 1 for surface rock masses with low stress.
B1 < 1 and B2 < 1 for tight highly stressed rock
masses
bh: base width of hill or width of valley [L]
Co : uniaxial compressive strength [M/T2L]
C: minimum rock cover [L]
Ceff : effective rock cover [L]
Cv : vertical rock cover [L]
Ch : horizontal rock cover [L]
Cov : vertical overburden cover [L]
Ccorr : remaining cover after topographic correction [L]
c’ : effective cohesion of sliding joint [M/T2L]
CAR: cover alteration ratio
D: disturbance of rock (between 0: undisturbed and 1: highly disturbed)
DB: borehole diameter [L]
DH: thickness of the decompressed zone (topographic correction models) [L]
DT: tunnel diameter [L]
d: distance of the structure from tunnel center [L]
dh: heigth of hill or depth of valley [L]
ei : hydraulic aperture of joint set i [L]
em : mechanical aperture [L]

ix
E1: deformation modulus of the rock mass in the models [M/T2L]
E2: deformation modulus of the structures in the models [M/T2L]
E’1: deformation modulus in the principal axis of anisotropy of the rock mass
in the anisotropic models [M/T2L]
E’2: deformation modulus the secondary axis of anisotropy of the structures in
the anisotropic models [M/T2L]
Erm: rock mass deformation modulus [M/T2L]
Eir: intact rock deformation modulus [M/T2L]
FS : factor of safety
H0 : natural water table head [L]
Hi : tunnel internal water head [L]
H: driving (effective) head inside the tunnel (or the test zone) [L]
i: hydraulic gradient [L/L]
JRC : joint roughness coefficient
JCS : joint compression strength [M/T2L]
K: stress ratio
K’: secondary stress ratio
Kw: bulk modulus of water [M/T2L]
kc : equivalent continuum permeability [L/T]
kd : discrete feature permeability [L/T]
k0 : initial rock mass permeability [L/T]
ki : permeability tensor for joint set i [L/T]
kp : physical permeability [L2]
l: surface of the sliding joint of unitary width [L2]
Lb: length of the tested borehole section [L]
Lc : length of the flow channel [L]
M: number of joint set
n: porosity [%]
Pc : critical water pressure to initiate jacking or shearing [M/T2L]
Pj : jacking water pressure [M/T2L]
Ps : shut-in water pressure [M/T2L]

x
P: water pressure [M/T2L]
Q: Q system rock mass classification
Qc : water flow through an equivalent continuum [L3/T]
Qd : water flow through a discrete structure [L3/T]
r: radius [L]
R: pressurized water reach (= 2 H0) [L]
RMR: rock mass rating [%]
si : joint spacing of joint set i [L]
S: ratio of contact area with the total area of the crack [L]
T: thickness of structural zones in the models [L]
U: uplift water pressure [ML/T2]
U1 ; U2 : resultant pressure force using non-linear distribution [ML/T2]
U1L ; U2L : resultant pressure force using linear distribution [ML/T2]
u: water pressure [M/T2L]
u(r): radial displacement [L]
u: displacement [L]
V: horizontal water pressure [ML/T2]
W: D-shape tunnel width [L]
Wb : weight of the sliding block [ML/T2]
Z: degree of disturbance due to blast damage (from 0 for undisturbed rock
masses to 1 for very disturbed rock masses)
z: vertical depth [L]

xi
CHAPTER 1

INTRODUCTION
1.1 Background

Pressure tunnels are largely used to convey water to hydroelectric powerhouses.


Usually, each project has a forebay reservoir, an intake, a pressurized tunnel, a
steel lined section connected to a surface or underground powerhouse, and finally
a tailrace tunnel or channel (fig. 1.1).

Unlined pressure tunnels were first used by the Norwegians in hard granitic rock
masses, which would constitute the ideal material for this type of structure.
However, the increase in the demand for electrical energy has pushed the industry
to build projects in sites with less favourable geology all over the world. In the
present study, the common definition of unlined pressure tunnels is everything but
steel lined. Figure 1.2 presents the different types of lining. An unlined pressure
tunnel is always pervious to a certain extent and if the rock is hard, insoluble and
non-porous and the rock mass is of good quality, unlined tunnels are usually
supported by rock bolts and local shotcrete. When the intact rock is slightly
subjected to dissolution, or is porous or with a poor rock mass quality, shotcrete
and concrete are used. Concrete can also be used to enhance the hydraulical
characteristics of the tunnel; however, because shotcrete and concrete will always
crack to some extent due to cold joints, shrinking and deformation, they cannot be
considered as a totally impervious liner.

When it is properly reinforced with steel, a concrete lined tunnel is considered to


be semi-pervious because of the controlled cracking that results from tension
forces (Benson, 1988). The reinforcement must be carefully designed according to
the elasticity of the rock mass and the in-situ stress, to limit cracking and
deformation. This type of lining provides a long term semi-pervious liner which
greatly contributes to water pressure losses through the controlled cracks. Slight
lack of minimum stress can also be treated by pre-stressing the rock mass with
high pressure grouting; however, this method is not discussed in this thesis.

1
Lined tunnels are commonly defined as a steel liner cast in concrete and
completed with some annulus grouting to fill voids around the steel liner system
(steel/concrete/rock mass). The steel can also be replaced by a synthetic
membrane, but in this case, it needs to be a sandwich liner because of the need to
support the membrane against external water pressure in case the tunnel is
dewatered. These types of lining are totally impervious and careful design must
be carried out in order to account for the main interactions, which are mostly
internal pressure and external pressure when the tunnel is dewatered. The
participation of the rock mass could also be taken into account, but it is not the
common practice. The length and location of lined sections are one of the key
design problems for pressure tunnel projects, because this type of lining is very
expensive to build.

2
4a 5
1

2 3

6a

9 10
δ 8 12

7 11

1
4b
2
5

6b 9 10
δ
8 12

7 11

6a
9
δ 8 12

7 11

1 – Upper reservoir 2 – Intake


3 – Headrace tunnel 4a – Shallow valley
4b – Deep valley 5 – Surge shaft
6a – Pressure shaft (55° ≤ δ ≤ 90°) 6b – Slightly inclines pressure tunnel ( δ ≤ 15°)
7 - Debris trap 8 – Steel lined pressure tunnel
9 – Underground powerhouse 10 – Surface powerhouse
11 – Tailrace tunnel or channel 12 – Lower reservoir

Figure 1.1 – Typical pressure tunnel and shaft


schemes for hydroelectric projects

3
Grout
Rock bolt

Steel Concrete Shotcrete

Concrete

STEEL LINED SEMI-LINED UNLINED


Steel liner embedded into Concrete liner reinforced Rockbolts and shotcrete
concrete and grouted or not, grouted

Figure 1.2 – Typical type of pressure tunnel and shaft lining


for hydroelectric projects

Thus the choice of the type and location of steel lining and concrete plugs is of the
highest importance in the design of pressure tunnels. Minimizing steel lined
sections and optimizing unlined tunnel sections are key to achieving an
economical project. Available rock cover criteria should be used with great care
for preliminary design, because it can over-estimate the minimum stress and thus
allow for hydro-jacking to occur if it is used in the final lining design. Second, it
addresses only a part of the problem. For example, it does not consider other
important parameters such as the rock mass quality, initial pore pressure
distribution, tunnel diameter and geometry (in the case of shallow tunnels), rock
mass permeability, etc. The use of cover criteria without sufficient knowledge of
these important parameters have lead to the design of unsatisfactory projects,
resulting in repair expenses and a loss of revenue. This might be largely due to the
use of conventional rock cover criteria, developed for massive and hard rock
masses, in areas where failure may be controlled by other parameters such as
erosion, weak rock mass, high permeability or adverse geological features and
topographies.

4
1.2 Study problem

The publication of the renowned empirical criteria by the Norwegians in the 70’s
and the 80’s, has established the foundation of the modern unlined pressure tunnel
preliminary design. Equation 1.1 and figure 1.3 illustrate the criterion as proposed
by Berg-Christensen and Dannevig (1971) with the cosine correction for the
slope.

FSγ w H
C≥ (1.1)
γr cos β

H
C
β

Figure 1.3 – Norwegian criterion (modified from Berg-Christensen


and Dannevig, 1971)

Equation 1.1 provides the minimum rock cover needed for the unlined section,
according to factors of safety, gravity stress and slope correction. Anywhere along
a tunnel, areas that present less cover than the minimum cover from equation 1.1
should be steel lined.

The criterion was later improved by Broch (1982) in order to make corrections for
topographic noses. It was assumed that these decompressed topographic features
would not contribute to the increase in minimum stress. Consequently, the
estimated cover should consider a trimmed topography. Figure 1.4 illustrates this
correction.

5
900

800
11
130

10
00
12

00
140

00
0
1500

0
1600
1800

1700

Concrete lined shaft

Unlined tunnel el
nn
Steel lined tu

Figure 1.4 – Topographic correction (modified from Broch,1982)

However, the drawbacks of such simple and easy to use criterion are the
generalization and the over simplification that can be done by non specialists,
leading to borderline tunnel safety design. They can over-estimate the minimum
stress and thus allow for hydro-jacking to occur if it is used in the final lining
design. It was acknowledged by several authors that gravity criteria, which
consider the weight of the overlying rock, were meant only for good massive
granitic rock masses and that they did not apply to weaker rock masses. In
addition, empirical criteria do not consider geological features and groundwater
conditions; the only controlling parameter being the topography.

This problematic issue was addressed by Norwegians engineers and they


recommended a comprehensive investigation into tunnel alignment, especially in-
situ stress measurements. With the developments of in-situ stress measurements, a
new design criterion was proposed (e.g. Brekke and Ripley, 1987). The steel liner
was extended until the minimum measures stress time a factor of safety was equal
to the internal tunnel static pressure.

6
But the simplicity of the initial empirical criteria remains deeply anchored within
the engineering community and the recommendations provided with the criterion
are sometimes overlooked.

Finally, the design of unlined pressure tunnels usually takes several years to be
completed; from the initial project statement to the construction of the project.
During this period, the project is in constant evolution and so is the design.
Moreover, unlined pressure tunnels are large civil structures, built in a complex
environment, sometimes covering several kilometres, with possible different
geologic/structural, topographic and hydrogeologic features along its axis.
Therefore, problems arising from unlined tunnel design are not simple and they
require careful study and analysis.

1.3 Research objective

The fundamental problem studied in this thesis is the preliminary design of safe
and economic unlined pressure tunnels. The main objectives are the following:

1. To study the robustness of the Norwegian criterion using numerical


modelling with a new cover alteration criterion;
2. To investigate different concepts that could alter the rock cover such as,
the cover versus tunnel diameter, topographies, structural features like
faults and dykes;
3. To propose new design guidelines for the estimation of rock cover at the
preliminary stage of a project based on numerical modelling results;
4. To validate the guidelines with case studies and existing
recommendations.

Computer models were developed in order to study the application limits of


preliminary cover criteria. The objective of numerical modelling is to make a
large number of simple stress/strain analyses with topographic features such as
hill and valley, and compare the minimum stress obtained from the numerical
model with the theoretical minimum stress given by the Norwegian criterion,

7
including the cosine correction and any topographic shaving. In addition, the
presence of structural features, such as softer or harder rock zones, is studied.

The statements behind largely used cover criteria are reinterpreted in order to
place their purpose at the preliminary stage of projects when little information is
available. Furthermore, new views are proposed for the development of fruitful
investigations as well as new relations for tunnel stability and leakage assessment.

Finally, based on the review of existing design recommendations, recent and


historical case studies, results from numerical modelling and design guidelines are
presented for preliminary layout, investigation orientation and planning, as well as
final design.

1.4 Thesis structure

The thesis reports on the following tasks:

1. Present a state of the art literature review;


2. Identify parameters controlling unlined pressure tunnel behaviour;
3. Perform numerical modelling and compare results with the Norwegian
criterion;
4. Discuss investigation objectives and recommended means;
5. Propose design guidelines based on study results.

Chapter 2 presents a state of the art literature review to identify ways of


improvement in the design practices. Chapter 3 discusses the parameters
controlling unlined pressure tunnel behaviour. The numerical modelling carried
out to investigate the Norwegian criterion is presented in Chapter 4.

Numerical modelling allows for the visualization of some important


topographical/structural stress alterations around the tunnel that could lead to the
overestimation of the minimum available confinement around a tunnel. This is the
most critical mechanism in this study i.e., stress alteration around the tunnel that
does not contribute to tunnel confinement because of large tunnel diameters, soft
zones, steep valley side, etc.

8
To a lesser extent, some models were carried out to study the water
pressure/permeability coupling interaction. The total relative leakage was taken as
the model output while varying the deformation parameters. This was done with
the help of the permeability versus volumetric strain relation from Lombardi
(1989a). The flow study models have allowed for the study of the potential effect
of rock mass quality and how it might mislead preliminary layout.

Verifications for the validity of the proposed guidelines were made using existing
known project performances. Chapter 5 presents the numerical modelling results
and identifies the basic observations. The proposed design guidelines and their
validation for preliminary design are presented in chapter 6. Chapter 7 discusses
the basic assumptions and limitations and emphasizes the importance of
engineering judgment. It also provides the principal conclusions of the study and
recommendations for future work.

9
CHAPTER 2

METHODS OF UNLINED PRESSURE TUNNEL DESIGN


2.1 Types of unlined pressure tunnel failure

There are two principal modes of failure for unlined pressure tunnels: excessive
leakage and geomechanical instabilities. Depending on the scale of the
phenomenon it can be either global failure or local failure.

2.1.1 Excessive leakage

Excessive leakage can be the result of two different phenomena:

1. high initial natural permeability;


2. high induced permeability due to hydraulic failure.

High initial permeability can be due either to the rock matrix, the rock mass, or
any locally untreated high permeability zone. On the other hand, high
permeability due to hydraulic failure is generally defined as hydraulic jacking,
which is described as the opening and/or shearing of existing joints and fissures
that can occur when the water pressure reduces effective principal stresses in such
a way that the system reaches failure envelope in the tunnel vicinity which results
in a large increase in permeability (Rancourt and Mitri, 2007).

For the purpose of evaluating the amount of water flowing in or out of a tunnel, a
high initial (natural) pore pressure will limit the total amount of flow (Deere,
1983) because it will counteract the internal tunnel pressure. In extreme cases, if
the initial pore pressure is higher than the pressure inside the tunnel, there can be
no outflow, but rather an inflow (Fernandez, 1994; Alvarez, 1997; Merritt, 1999,
Rancourt et al., 2006b). However, near a powerhouse or a plug, or in the case of a
shallow tunnel, where atmospheric pressure is approaching the pressurized zone
resulting in a high local gradient, the initial pore pressure may be locally drained
and, therefore, not help to reduce the outflow. Conventional design usually
neglects the effect of natural pore pressure, because it is difficult to measure

10
accurately and it places the design on a conservative side. This aspect will be
discussed further in chapter 3.

Merritt (1999) addressed the subject of total tunnel leakage and considers leakage
rates of about 5 l/s/km to be acceptable for most projects. The same author
suggests that leakage exceeding 10 l/s/km might be unacceptable. However, these
proposed values might differ from project to project, depending on the
economical value of water and other sociological impacts. As described in Benson
(1988), leakage can have serious consequences, other than excessive water loss,
such as the flow of high pressure water into nearby openings, potential slope
instability and permanent rock mass deformations.

2.1.1.1 High natural permeability

Excessive leakage can be caused by highly initial permeable matrix rock or


geological features that have not been properly consolidated, treated and/or
grouted during construction. Leakage problems with initially permeable rock
masses have been observed even in areas with sufficient in-situ stresses (Brekke
and Ripley, 1987). Leakage caused by natural rock mass permeability can be
estimated using the Darcy law if the rock mass can be considered as an equivalent
permeability continuum, provided that the studied representative volume is large
enough (Long et al., 1982). This is usually the case for pressure tunnels.

Equivalent continuum

A definition given by Long et al. (1982) was that fluid flow can be considered to
occur as an equivalent permeability continuum under a considered scale, only if
there is an insignificant change in the value of the equivalent permeability with a
small addition or subtraction to the test volume and an equivalent permeability
tensor exists. This predicts the correct flux when the direction of a constant
gradient is changed. This subject will be examined further in chapter 3. When
these criteria are met, the Darcy law can be used and is expressed as:

Qc = kciA (2.1)

11
Where
Qc : the continuum water flow [L3/T]
kc : continuum permeability [L/T]
i : hydraulic gradient [L/L]
A : flow cross-sectional area [L2]

Methods for estimating the rate of outflow were originally addressed by Bouvard
and Pinto (1969), and by Schleiss (1986). Their solution considers an undefined
reach of flow at which the seepage becomes negligible. Fernandez (1994)
proposed that the reach of flow would be given by the location of the original
water table. Therefore the rate of inflow or outflow of a circular pressure tunnel
per unit of length, can be estimated by:

Q =
(
2π kc H i − H 0 )= 2π kc (H )
(2.2)
c
R R
ln ln
r r

This relation is quite similar to the one given by Goodman et al. (1965) for the
estimation of water inflow. In equation 2.2, kc is the equivalent continuum rock
mass permeability, Hi internal water pressure head in the tunnel, H0 elevation of
the original water table head relative to tunnel, R = 2 H0 and r the tunnel radius.
Equation 2.2 was used by Rancourt et al. (2006b) to successfully estimate the total
expected leakage for the Toulnustouc 10 km long pressure tunnel located in the
north coast of Quebec.

Discrete continuum

When the media cannot be considered as an equivalent continuum, the theory of


fluid flow in rock joints is based on laminar flow between a parallel pair of
smooth plates which is the base of the cubic law. Several authors have worked on
the subject, starting with Poiseuille (1844) and Boussinesq (1868), followed by
Lomize (1951), Polubarinova-Kochina (1962), Snow (1965), Romm (1966) and
Louis (1969) amongst many others. However, Snow (1968), and later

12
Witherspoon et al. (1979), gave the following well-adopted relation called the
cubic law:

γ w e i3
Qd = (H i − H 0 ) (2.3)
12 μ L c

with Qd : discrete water flow [L3/T]


Lc : Length of the flow channel [L]
ei : the hydraulic joint aperture [L]
γw , μ : the unit weight of water and the dynamic viscosity of water
respectively (9.81 kN/m3), (10-3 N•sec/m2)

Equivalent continuum versus discrete continuum

Both concepts of equivalent continuum permeability and discrete permeability


have to be addressed in the leakage estimation. Few discrete geological features
can control the majority of the leakage flow for a given project. Consequently,
equivalent continuum can be assumed initially, but the investigation program
should be oriented towards the finding of discrete permeable features along the
tunnel.

The study of water flow, in fissured mediums, needs to address the problem of
scale in order to be able to choose which approach is required, the discrete
discontinuous media or the equivalent continuum media, or a combination of
both. Figure 2.1 illustrates the differences between both approaches; case 1 and
case 2 represents an equivalent continuum medium, and case 3 and case 4 clearly
apply to discontinuous media.

13
1 Intact 2 Heavily jointed

Jointed Blocky
3 4

Figure 2.1 – Schematic showing difference between various rock media

2.1.1.2 Hydraulic failure

Excessive leakage can also be the consequence of high induced permeability due
to joint opening or shearing under high water pressure. The broad notion of
hydraulic failure includes at least two failure modes: hydraulic jacking and
hydraulic shearing. Hydraulic jacking was basically defined by the Norwegians
(e.g. Broch, 1984) to be that nowhere along the unlined pressure tunnel or shaft
should the internal water pressure exceed the minor in-situ principal stress. This
criterion is general and assumes that at least one joint perpendicular to the
minimum principal stress exists. It also does not take into account the possible
anisotropy of stress leading to high shear stress.

When a joint is actually oriented perpendicular to the minimum principal stress,


the jacking is defined when the internal water pressure equals the total stress
acting perpendicular to the joint, thus rendering the effective stress to zero. In
these stress conditions, the joint aperture is then controlled by movements and
deformations of the surrounding rock blocks and the joint is in jacked condition.
Last and Harper (1989) gave the following definition for pure hydraulic jacking:

14
σ 'j = 0 (2.4)

with
σ’j : the effective stress perpendicular to the jacked joint

It holds that if the resistance of the rock mass follows a Coulomb type law, then
pore pressure can be integrated directly into the following well-known shear
criterion written in principal stress:

⎛ φ⎞
σ 1 = C0 + σ 3 tan 2 ⎜ 45 + ⎟ (2.5)
⎝ 2⎠

Equation 2.5 can be written in effective stress, and become:

⎛ φ' ⎞
(σ 1 − u ) = C0 + (σ 3 − u ) tan 2 ⎜ 45 + ⎟ (2.6)
⎝ 2⎠

Then we can isolate u to figure out the pressure required to initiate Coulomb
failure into the rock mass for a certain stress state. We have the following
equation:

(σ 1 − σ 3 ) − C0
u = σ3 − (2.7)
⎛ φ' ⎞
tan 2 ⎜ 45 + ⎟ − 1
⎝ 2⎠

As illustrated in equation 2.7, if we keep σ3 constant, the increase of σ1 will


decrease the critical failure pressure. This shows that even if σ3 is high enough to
withstand hydraulic jacking, with a high σ1 rock mass failure by shearing is still
possible with high stress anisotropy.

This is supported by Barton (1986) who addressed the fact that even with
sufficient in-situ minimum stress, some hydroelectric projects have suffered
leakage immediately at the upstream of the steel liners due to transfer of strain to
the rock mass and related with joint shearing. This phenomenon was also
observed by Dalho et al (2003), who suggested that the large water flow observed
in hydro-jacking tests could be related to fracture shearing.

15
These observations could be partly explained by the fact that permeability is
greatly increased when the peak shear strength is exceeded and dilation occurs.
Olsson and Brown (1993) noted that shear displacement causes an order of
magnitude increase in permeability compared to 2 to 3 for hydraulic jacking.
These results illustrate the importance of minimum stress and stress anisotropy on
tunnel stability against hydraulic failure. Chapter 3 will address some important
aspects of the in-situ state of stress around the tunnel in greater detail.

2.1.2 Geomechanical instabilities

This section includes both underground and surface instabilities. It is crucial that
the tunnel remains open for the life of the project. In addition, tunnel operation
might have unfavourable influence on the surrounding area, especially the
overburden located on surface.

2.1.2.1 Tunnel collapse

One of the major causes of unlined pressure tunnel collapse is the rock matrix
dissolution and/or erosion of joint fillings. Tunnel zones subject to such
difficulties must be treated with appropriate long term consolidation in order to
remove the risk of collapse. Furthermore, long term structural stability must be
assured with the use of appropriate consolidation preventing wedge loosening.
This aspect must take into account the operational requirements of the tunnel
(dewatering, dynamic effects, etc.). This topic can be separated into two distinct
operations: temporary ground support and permanent consolidation.

Ground support is meant for general short term tunnel stability (especially during
construction) and is also useful for long term operation. Ground support is
generally made of local or pattern rock bolting, with or without shotcrete
reinforcement. If the rock quality is poor, steel ribs may be required.

For construction safety, a flexible wire mesh is usually installed on the crown and
removed at the end of construction, in order to avoid friction head losses and
problems related to the fall of wire mesh on the invert and the rapid filling of the

16
debris trap. The flexible wire mesh can be left in place provided it was designed
for long term performance.

Consolidation is usually considered for long term tunnel operation. It is


particularly needed for erodible rock mass, poor general quality and highly
fractured zones. Consolidation is therefore defined as the final treatment that is
required considering long term conditions such as water flow, tunnel fill-up and
dewatering, and their effect on tunnel stability.

Consolidation requirements can vary greatly but usually include and complete the
ground support installed during tunnel excavation. As a result, consolidation is
made with grouted rock blots, steel ribs, reinforced shotcrete, with or without
concrete reinforcement, rock grouting, etc. Regarding the design of consolidation,
the deformability of the rock mass must be addressed because of its impact on the
required reinforcement to control cracking distribution. The topic of tunnel
consolidation is a large one and it is not the objective of this study to give a
complete description of the subject. The reader is directed toward classic books
such as Wahlstrom (1973) for general tunnel design, or Hoek et al. (1995) for
hard rock tunnelling, and Einstein (1991) for soft rock tunnelling.

2.1.2.2 Surface instability

This failure mode is a mix of jacking and shearing, along with the condition that
the ground surface is close enough and that major joints exist and are parallel to
the ground surface (Ming and Brown, 1988; Alvarez, 1997). The problem of
hydraulic pressure acting on rock slopes was initially addressed by
Terzaghi (1963), who provided the first expression for the slipping of a rock
wedge along existing bedding planes filled with water. The expression was further
developed by Hoek and Bray (1974) who gave to the following equation:

c' l + (Wb cos θ − U − V sin θ ) tan φ '


FS = (2.8)
W sin θ + V cos θ
with
FS : Factor of safety against sliding of block

17
c’ : Effective cohesion of sliding joint [M/T2L]
l : Surface of the sliding joint of unitary width [L2]
Wb : Weight of the sliding block [ML/T2]
U : Uplift water pressure [ML/T2]
V : Horizontal water pressure [ML/T2]
θ : Angle of the sliding surface according with horizontal [angle]
φ’ : Effective friction angle of the sliding joint [angle]

Figure 2.2 illustrates the slope stability analysis of a rock wedge subjected to
water pressure. Figure 2.2 shows pore pressure distribution as linear, however the
irregular nature of pore pressure and the low porosities of rock masses make the
linearity assumption questionable. It will be shown in the next section that
Alvarez (1997) proposed a non linear pore pressure assumption.

Tension crack

W
Sliding joint

Q
U
l

Figure 2.2 – Slope stability analysis of a rock wedge subjected to water


pressure (modified from Hoek and Bray, 1974)

Benson (1987) revealed that if water flows towards the surface, it might
destabilize the overburden and possibly induce landslides. However, the study of
landslides of soils caused by pressure tunnel leakage is beyond the scope of this
study; those interested should refer to Abramson et al. (2002) for further reading.

18
2.2 Existing design criteria

The purpose of the following sections is to present a review of the design criteria
that are available in the literature. Design criteria have mostly been developed and
updated following unsuccessful projects and costly remedial work and are,
therefore, basically empirical in nature.

Historical design criteria have been widely based on rock cover assumption, and
were subsequently adjusted to account for topographic effects and structural
features. Table 2.1 presents an overview of the evolution of the design criteria.
According to table 2.1, the first pressure tunnel criterion was given by
Talobre (1954), who stipulates that the topographic level above the tunnel should
always be equal or greater than the reservoir water level. This highly conservative
criterion was based on mostly carbonate, argillaceous and arenaceous sedimentary
rocks found in France. Terzaghi (1963) later proposed that the expression of
vertical stress has the weight of the overburden vertical column and assumed that
vertical stress corresponds with minor principal stress. Snowy Mountain
Hydroelectric Authorities in Australia (Dann et al. 1964), proposed that
everywhere along the tunnel, the side cover would be twice the length of the
vertical cover. Berg-Christensen and Dannevig (1971) presented the initial
Norwegian criterion derived from Terzaghi and with a correction for slope
proximity. In the early 80’s, Broch (1982) proposed a correction for topographic
noses. The assumption was that these topographic features were decompressed
and did not contribute to stress deeper in the rock mass. Finally, Alvarez (1997)
presented the Illinois criterion which basically accounts for structural features in
close proximity to the slope. The criterion from Alvarez (1997) is mostly based
on the work from Hoek and Bray (1974) with non linear water pressure
distribution along the joints.

19
Table 2.1 – Evolution of cover criteria

Method Period Relation Remarks References


Topographic 50’s and C≥H -For an unlined tunnel, Talobre (1954, 1967)
level 60’s the internal pressure
must be below ground
H level
C

Vertical 50’s and FSγ w H -The origin of cover Terzaghi (1962)


weight of rock earlier C≥ criterion assuming the
γr vertical stress is the
minor principal stress

H
C

20
Table 2.1 – Evolution of cover criteria (continued)

Method Period Relation Remarks References


Snowy 50’s and FSγ w H -Consider side cover Dann et al.
Mountains 60’s C≥ (1964)
γr
Ch = 2Cv
H
Cv
C

0.2Cv

2Cv

Norwegian’s 70’s FSγ w H -Correction for Berg-


C≥ slope proximity Christensen and
γ r cos β Dannevig
(1971)

H
C
β

21
Table 2.1 – Evolution of cover criteria (continued)

Method Period Relation Remarks References


Modified 80’s FSγ w H -Topographic Broch (1982,
Norwegian’s C≥ correction 1984)
γr cos β

900

800
11
130

10
00
12

00
140

00
0
1500

0
1600
1800

1700
Concrete lined shaft
l
Unlined tunne
nel
Steel lined tun

Modified 80’s FSγ w H


Norwegian’s C≥ -Cross section
γr cos β

H
CV

22
Table 2.1 – Evolution of cover criteria (continued)

Method Period Relation Remarks References


Illinois 90’s C =γ w M -Includes structural Alvarez
University1 H γr λ features with slope (1997)
β θ1 = λ − β
proximity
θ1 θ2 = β − α

H 1
W Hydrojacking joint set

C
U1
α
2 θ2

H
B1 = U1 / U1L H
B2 = U2 / U2L

Modified 90’s -Includes the weight Eskilsson


Norwegian’s C = 0.5 H of overlying (1999)
Special case overburden
for Cov -a special case of
flat the Norwegian
H
topography criterion with β = 0
with Cv and FS = 1.3
overburden
influence 2

1. M =
cos β (A − K + C ) where
D
A= (B1 − B2)⎛⎜

tan β
tan λ
− tan α ⎞⎟ ; K = tan α ⎛⎜ B1 + B2
tan( φ 2 −α ) ⎠ tan λ ⎝
tan β ⎞ ;
tan( φ 2 −α ) ⎠


tan β ⎞

(
tan α
⎟ C = ⎜ B2 + B1 tan( φ 2 −α ) ⎟ ; D = 1− tan λ )
2. C = C ov γ ov + C v γ r

23
2.3 Design methods

Design methods address more global approaches to the design, taking into
account not only the weight of the overburden, but also the measured in-situ
stresses, tunnel geology, the use of instrumentation, etc. The Norwegians first
proposed methods to account for several basic parameters other than only the
available rock cover (Broch, 1982; Brekke and Ripley, 1987). The method given
by Merritt (1999) is the most complete tool available to date. Merritt (1999)
presents a design chart that includes topographic situations, pre-excavation natural
pore pressure, measured in-situ minimum stresses and rock permeability related to
internal tunnel pressure. Table 2.2 presents an overview of the evolution of some
of the design methodologies presented over the years. Figure 2.3 illustrates
Merritt’s methodology (1999).

24
Table 2.2 - Evolution of design methodologies
Method Period Features Assumptions References
Jaeger 60’s Analytical solution of failure conditions - The rock is homogeneous and Jaeger (1961)
for: isotropic
1. Hard unfissured rock - The rock mass is infinitely large
2. Radially fissured rock - The tensile circumferential stress
3. Plastic rock induced by water pressure is
4. Tunnel near a rock slope proportional to R/x
(R the tunnel radius, x the distance
from the tunnel centre)
- No water in the rock mass
Talobre 50’s – -No liner if water head is lower than Talobre (1967)
60’s 1. Good geological knowledge ground thickness above tunnel (C≥H)
2. Applies to sedimentary rock -If C is between H and 0.6H, the seal
masses is not guaranteed
3. When C≤H, water pressure -At C between 0.6H and 0.4H, there
tests are required. is risk of piping of loose filling
material
-For C < 0.4H steel liner is required
Hoek 80’s Analytical solution for condition of - The tunnel modifies the stresses Hoek (1982)
joints opening in plane strain according to the Kirsch equations
1. Cover criterion for flat ground - Plane strain conditions
and K = σh / σv > 1 - K has great influence on jacking
2. In-situ pressure tests conditions
3. Geological mapping during - The tunnel direction is assumed to
construction be parallel to one principal stress
directions
Norwegian 80’s Developed from the cover criterion, but - Nowhere along the pressure tunnel Broch (1982);
expressed in terms of in-situ stresses or shaft should the internal water Brekke and
1. Perform stress measurements pressure exceed the minor in-situ Ripley (1987);
2. Geological mapping during principal stress Benson (1987)
construction - Massive hard granite and gneiss
3. Monitoring during
construction
Ming 80’s Numerical study of structural failures -Mainly based on Hoek (1982) and Ming (1987),
1. Present relation for hydraulic using Barton (1986) joint shear Ming and
jacking of vertical and strength relation Brown (1988)
horizontal joints
2. Relation for hydraulic
shearing and the importance
of K
Merritt 90’s Overview of geologic and geotechnical -Most complete to date Merritt (1999)
considerations: -Addresses rock dissolution
1. includes topographic -Based on the Norwegian’s minimum
situations stress approach
2. natural pore pressure -If natural pore pressure in the rock
3. measured in-situ minimum mass is higher than the pressure in
stresses the tunnel, there is no leakage
4. rock permeability

Table 2.2 reveals that in 1961, Jaeger was the first to analytically study the failure
behaviour of pressurized rock tunnels. During the same period Talobre (1967),
pushed his criterion further, acknowledging that for good quality rock mass, the
cover could be reduced to some extent from the initial conservative criterion
given in table 2.1. In the 80’s Hoek (1982), continued the work from Jaeger
(1961), providing analytical solutions based on Kirsch’s equations. This type of

25
approach was also used by Ming (1987). The Norwegians, with the work of Broch
(1982) and Brekke and Ripley (1987), gave a description of the limits of the cover
criterion and recommended to perform in-situ stress measurements, to consider
the geology and to do performance monitoring. Merritt (1999), expanded this
method to include natural pore pressure acting against water pressure and also
permeability and the dissolution of rock masses.

H0 Ground surface Surge Shaft

Hi
UPT
C B (A?) C B A

Underground Powerhouse

LINING SELECTION
ROCK
STEEL
SOLUBLE
LEAKAGE HIGH Erm
GROUT, CRACK CONTROL
HIGH
ZONE C ZONE B
TY

NON- MOD. Erm REINFORCED CONCRETE,


LI

SOLUBLE MEMBRANE
BI

CONTROL INFLOW
LOW Erm
EA

STEEL, MEMBRANE
RM
R

SEAL CRACKS
O LUB
SO

PE
C
K LE

GROUT ROCK
H

STEEL, MEMBRANE
G

SOLUBLE
HI

Hi < H0 Hi > H0 MED. PERM. LEAKAGE HIGH Erm


γwH < σmin GROUT, SEAL CRACK
INFLOW OUTFLOW MED. NON- MOD. Erm REINFORCED CONCRETE
LO
E
BL

SOLUBLE
W
LU

LOW Erm REINFORCED CONCRETE


PE
SO
N- K

R
NO OC

NONE
EA
R

ROCK
BI

REQUIRED CONCRETE LINING, CRACK CONTROL


LI

SOLUBLE
TY

LEAKAGE HIGH Erm


NO LINING
Hi = Internal water pressure
ZONE A LOW NON- MOD. Erm
σmin = Minimum in-situ rock stress γwH > σmin Steel lining SOLUBLE CONCRETE LINING, CRACK CONTROL
LOW Erm
H = Hydraulic driving head REINFORCED CONCRETE
H0 = Natural pore pressure
Erm = In-situ rock mass modulus

Figure 2.3 - Unlined pressure tunnel design methodology


(modified from Merritt, 1999)

The starting point of figure 2.3 is the minimum stress criterion briefly discussed in
section 2.1.1.2, which requires that nowhere along the unlined pressure tunnel or
shaft should the internal driving water pressure (γwH) exceed the minor in-situ
principal stress (σmin). This criterion implies, however, that minimum stress was
measured in the field. With the aforementioned measurements available, the first
step is to determine in which zone (A, B or C) the project is located. For zone A,
the minimum in-situ stress is lower than the driving pressure (γwH) and the tunnel
must be steel lined. Zones B and C require that the minimum stress be greater
than the tunnel driving pressure (γwH) and, therefore, no hydraulic jacking is

26
possible. If the tunnel driving pressure (γwH) is above zero there will be outflow
(zone B). Conversely, if the tunnel driving pressure is negative, there will be
inflow (zone C).

The next step for zone B involves the permeability and the likelihood of rock
dissolution. For zone C, only rock dissolution needs to be verified.

This method assumes that in-situ stress is known and that the project has been
subjected to careful site investigation. The present study attempts to study the
limits of applicability of preliminary cover criteria at the beginning of the project,
before site investigation takes place.

27
CHAPTER 3

PARAMETERS CONTROLLING UNLINED PRESSURE TUNNEL


BEHAVIOUR
3.1 Foreword

Since the mechanisms of pressure tunnels are very complex and include several
parameters and interactions, it is helpful to start the analysis with the
identification of the principal parameters based on a general classification,
initially presented by Hudson (1992), called the rock engineering system. Later,
Jiao (1995) formalized this system and studied the particular case of an unlined
pressure tunnel. Nine principal variables were used in the study: in-situ stress,
induced stress and discontinuity intensity, shear strength of discontinuities, water
pressure, water leakage, displacement, as well as joint aperture and tunnel
dimensions. His results identify the critical interactions between in-situ stress and
water leakage, water pressure, water leakage, as well as between aperture and
water leakage.

The present study uses the results of Jiao (1995) as a starting point for design
method improvement. Parameters were divided into five basic groups: rock mass
geology, rock mass hydrogeology, in-situ stress, rock mass permeability, and
structural geology. For each group, investigation methods will also be discussed.
The end of this section will address the basic interactions between these five basic
parameters.

3.2 Rock mass quality

The rock mass quality is of vital importance to the design, construction and long
term operation of all types of tunnels. It is generally responsible for structural
stability problems encountered in tunnels. However, because of its intrinsic
complexity, it requires considerable attention in order to characterize
heterogeneity, anisotropy, non-linearity and identify associated uncertainties
(Hudson et al., 2001).

28
Rock mass includes discrete joint mechanisms and intact rock mechanisms into a
global rock mass mechanism. For the mean of unlined pressure tunnel design, two
kinds of characteristics are required: chemical dissolution and mechanical
properties.

Erosion and dissolution

Erodible and soluble rocks will likely impose a semi-lined section, therefore
precluding the used of unlined sections. Table 3.1 presents three different
mechanisms: erosion, alteration and dissolution. Table 3.1 also gives typical rock
type examples and the methods to use to test erodibility and solubility.

Table 3.1 - Mechanisms of erosion and dissolution

Physical Mechanism Behaviour Parameters Typical rock type References


phenomenon and tests
Erodability Mechanical Erosion Rock mass Poorly cemented Annandale
erosion quality and epiclastic sediments (2006)
fracture Fault and shear zone
orientation
Physical/chemical Slaking Slaking index Shale/argillite Franklin and
alteration and Pyroclastic rocks Chandra (1972)
weathering Smectite/illite/chlorite
Solubility Chemical Dissolution Solubility Halite James and
dissolution Solution rate Anhydrite Kirkpatrick
constant Gypsum (1980)
Carbonate rocks

Erodability is defined as a mechanical phenomenon that includes mechanical


erosion and erosion of slaking prone rocks. Mechanical erosion depends on the
rock mass characteristics, while slaking is mostly based on the presence of
swelling minerals. Van Schalkwyk et al. (1994) and Annandale (2006) have
addressed the erosion of rock masses located downstream from spillways. The
same basic principles can be applied for the assessment of unlined tunnel erosion.

Slaking behaviour should be considered separately because it exacerbates erosion


intensity. Extensive field and laboratory testing is required if the presence of such
weak rocks is possible along the tunnel alignment. The slaking property of a rock

29
is addressed with the slake durability test (ASTM D4644), originally proposed by
Franklin and Chandra (1972). Slaking is a characteristic of rocks containing
smectite, illite and chlorite clay-type minerals. These minerals are generally found
in argillaceous sedimentary rocks, pyroclastic rocks, and some ultramafic
intrusive rocks.

The solubility of rocks is also of great importance and is essentially a chemical


process. James and Kirkpatrick (1980) have proposed guidelines for the design of
dam foundations. The most soluble minerals are listed in decreasing order of
solubility: halite, anhydrite, gypsum and limestone. Limestone (Calcium
carbonate) is much more tolerant than the other minerals, which would require an
adequate semi-lined section.

Mechanical properties

Rock mass mechanical properties can be estimated using rock mass


classifications. These empirical classification guidelines are the Q system (Barton,
1974) and the GSI (Hoek, 1994), which replaced the former RMR (Bieniawski,
1973). The engineer using these classifications can advantageously divide the
rock mass into structural regions (rock mass partitioning or geomechanical
zoning) that are assumed to have the same properties and behaviour, with specific
support and consolidation requirements. In order to achieve the zoning and the
classification, it is essential that a consistent mapping be done along the tunnel,
describing all relevant features such as rock type, joint families, persistence,
spacing, roughness and aperture, soluble seams or zones, etc. Support and
consolidation requirements can then be examined.

Moreover, another very important parameter associated with rock mass quality is
the deformation characteristics of the rock mass. Rock mass deformability is
difficult to model and expensive in-situ measurements are required to acquire
knowledge of stiffness parameters, especially in the presence of pressurized water
that will cause a coupling effect between the rock mass deformation and the rock

30
mass porosity/permeability. Brekke and Ripley (1987) reported that several tunnel
failures have been associated with highly deformable geological structures.

Bieniawski (1978) examined the data from several projects where in-situ rock
deformability tests were carried out. He was the first to propose a graphical
relation between the RMR and the in-situ modulus of deformation. This study has
been followed by other researchers who have proposed different deformation
models. Table 3.2 summarizes some of the rock mass deformability models from
the literature. A more complete list of models is given in Hoek and Diederichs
(2006).

Table 3.2 Some available rock mass deformation models from the literature

Rock mass deformation models References


( RMR −10 ) Serafim and Pereira (1983)
Erm =10 40

E rm = Eir Rm Nicholson and Bieniawski (1990)

1 ⎛⎜ ⎞
RMR
22.82 ⎟
Rm = 0.0028 RMR 2
+ 0.9
100 ⎜⎝ ⎟

Erm =25log10 Q Grimstad and Barton (1993)

Eir ⎡ RMR ⎤ Mitri et al. (1994)


E rm =
2 ⎢⎣1 − cos(π 100 )⎥⎦

( )
Erm = 1− D C0 ⋅10
2 100
( RMR −10 )
40
Hoek and Brown (1997)

⎡ 1+ D / 2 ⎤ Hoek and Diederichs (2006)


Erm = Ei (0.02 + ⎢ (( 60 +15 D − GSI ) / 11) ⎥
⎣1 + e ⎦

Figure 3.1 shows the relation between the RMR and the predicted rock mass
modulus using some of the deformability models presented in table 3.2, using
Eir = 100 GPa, C0 = 90 MPa and D = 0. All the models, except one from Mitri et
al. (1994), which is a cosine function, are of the exponential type. For models
from Serafim and Pereira (1983) and from Hoek and Brown (1997), when RMR
values are above 70 %, the modulus becomes overly high. While other models

31
from Nicholson and Bieniawski (1990) and Mitri et al. (1994), consider the intact
rock modulus (Eir) as an input parameter and, therefore, converge toward Eir when
RMR = 100 %, which makes these models more physically realistic. More
recently, Hoek and Diederichs (2006) introduced a new database from China that
largely increased the number of data points. The distribution of GSI values
according to Erm from this new database is clearly of the exponential type. This is
why the model from Nicholson and Bieniawski (1990) was used in the modelling
presented in chapter 4.

200.0

180.0

Serafim and Pereira (1983)


160.0
Mitri et al. (1994)

140.0 Hoek and Brown (1997)

Nicholson and Bieniawski (1990)

120.0
Hoek and Diederichs (2006)
Erm (GPa)

100.0

80.0

60.0

40.0

20.0

0.0
0 20 40 60 80 100 120
RMR (%)

Figure 3.1 – Comparison of deformability models

In addition, important structures such as geological contact, faults, folds and shear
zones must be characterized and investigated separately because they might
control the failure behaviour of tunnel sections. This topic will be discussed
further in section 3.6.

3.2.1 Preliminary geomechanical assessment

At the preliminary stage of a project, the designer has generally very little
regional information from government agencies. At this early stage, possible
layout and tunnel alignments are studied from an economical standpoint, in order

32
to select the most cost-effective construction layout. Alignments at this stage are
not necessarily selected for their good geology.

Preliminary geomechanical assessments are therefore made based on regional


geology and past experiences in the area. The presence of large regional structures
such as faults and folds should also be regarded in order to safeguard against the
selection of potential problem sites.

The type of overburden might also be of interest, especially on hill slopes where
landslides have occurred in the past. The type of surface weathering is also
important; if the area presents a lateritic profile or has been subjected to the latest
glaciations, it might have an influence on the alignment and tunnel layout. Buried
valleys are always a concern since their depth is unknown, so careful investigation
should be planned for every valley crossing along the alignments. At the
preliminary stage, the most important rock parameter is the density to be used in
the Norwegian criterion. Table 3.3 presents the recommended ranges of density
from the general textbook literature.

Table 3.3 - Textbook density ranges according to rock type


Rock type Class Density Range Example
Igneous Intrusive 2.6 – 3.3 Granite, gabbro
Extrusive 2.4 – 3.0 Tuf, basalt, rhyolite
Metamorphic Non foliated 2.4 – 2.8 Quartzite, marble
Foliated 2.5 – 2.9 Gneiss, schists, slate
Sedimentary Clastic 2.2 – 2.8 Sandstone, shale
Organic and chemical 1.1 – 2.8 Coal, chalk, salt, gypsum, anhydrite
Carbonate 2.3 – 2.9 Limestone, dolomite

3.2.2 Rock mass investigation

After the preliminary stage, the project generally goes through on-site
geotechnical investigations to acquire better knowledge of local geology and
structures. At this stage, the designer is able to proceed with a detailed surface
mapping, rock mechanic classifications and geomechanical partitioning of the

33
tunnel. Sufficient work has to be carried out for each geological zone. Soluble
rocks, if present, have to be carefully studied.

The study of investigation results usually leads to more focused investigations to


study some aspect and/or to understand a particular feature in greater detail which
can add to the information gathered from previous investigations. Usually, the
investigation is carried out using surface geological mapping, geophysical survey,
boreholes, etc. Conventional laboratory tests as well as in place tests are routinely
done in order to allow rock mass classification according to Barton et al. (1974).

3.3 Hydrogeology

This section addresses the natural pore pressure distribution around the tunnel and
thus, the effective hydraulic pressure driving in or out of the tunnel. This pressure
results from the action of the internal tunnel pressure together with the natural
rock mass pore pressure that was initially acting along the tunnel vicinity. If, for
example, a Darcy-type material, the natural initial pore pressure is equal to the
tunnel pressure, there should be no outward leakage or inflow inside the tunnel,
because the hydraulic pressures will balance. In evaluating the amount of water
flowing in or out of the tunnel, a high initial (natural) pore pressure will limit the
total amount of flow out of the tunnel (Deere, 1983). In the extreme case, if the
initial pore pressure is higher than the pressure inside the tunnel, there can be no
outflow but rather inflow (Fernandez, 1994; Alvarez, 1997; Merritt, 1999).
However, near a powerhouse, a plug or in the case of a shallow tunnel, where
atmospheric pressure is approaching the pressurized zone resulting in a high local
gradient, the initial pore pressure may be locally drained and, consequently, not
help to reduce the outflow.

The following is based on the Terzaghi effective stress principle for soils. In the
case of water in a rock joint, the effective stress varies according to the following
relation (Cappa et al., 2005):

σ ' = σ tot − (1 − S ) P 0<S<1 (3.1)

34
with

S=Ac/A
Ac : Contact areas inside the crack [L2]
A : Total crack area [L2]

In this expression, S represents the ratio of contact area with the total area of the
crack. So, when S = 0, the joint is represented by parallel plates or is mechanically
open and when S = 1, water cannot develop pressure inside the joint, because the
joint is mechanically closed (Lombardi, 1989a). Therefore, it is obvious that S is
related to in-situ stress, rock type and rock mass structure. The effective stress
also interacts with joint deformation and, therefore, aperture, in a non-linear
fashion (Bandis et al., 1983), thus having a non-linear influence on water flow.
Table 3.4 presents a summary of some S values taken from the literature.

Table 3.4 - Summary of S values from the literature


Authors S Remarks
Rutqvist et al., 1998 0 Assumed
Cornet et al., 2003 0.25 Measured
Dalho et al., 2003 0.3 Assumed
Zangerl et al., 2003 0 Assumed
Cappa et al., 2005 0; 0.1; 0.2 Assumed
Bart et al., 2004 0.18; 0.22; 0.3 Marble
0.2; 0.23 Granite
0.24; 0.52 Schist
Asgian, 1989 0.2 Assumed
Noorishad et al., 1992 0 Assumed
Walsh, 1981 0.1 Polish granite surface
0.44 Tension fracture, σ’=0
Pine and Batchelor, 1984 0 Brittle rocks
Millard et al., 1995 0 Assumed
Lombardi, 1989a,b 0 Open fracture
1 Completely closed fracture

Generally, with most investigators working with brittle rocks, the value of S is
assumed to be zero (Terzaghi and Peck, 1967; Hoek and Bray, 1974). For a rock

35
mass, the actual pore pressure spatial distribution and, consequently, the spatial
distribution of S, is generally difficult to establish mainly because of the
occurrence of high permeability zones and open joints among unfractured and
probably unsaturated zones that are not connected to adjacent saturated zones.
Furthermore, joints with very small apertures might present capillary effects that
change the flow behaviour depending on the scale. Thus, the term groundwater
table is not very appropriate for rock masses, as it is closely related to the scale of
the observed area. In some rock mass types, it is difficult to accurately define the
water table, especially in hard granite or igneous rock with tight and widely
spaced discontinuous joints (Merritt, 1999).

This is why it is so important to characterize the natural pore pressures that are
effective around a pressure tunnel as much as possible. This aspect is often
neglected during investigation and design. Sometimes because of the lack of
quality data, the design is often done with the pore pressure designated as non
existent, thus setting conservative parameters for the design may be also because
of the fact that rock masses in tunnels are often dry over a wide area, with only a
few sparsely located open joints which provide water. In the field rock masses
might not be totally saturated because of the poor interconnection between
fractures. This observation was made by Norton and Knapp (1977) when
investigating field porosity measurements. They found great differences between
total porosity and flow porosity leading to the conclusion that flow porosities are
generally a small fraction of the total number of existing porosities. Parameters
like in-situ stress, depth and actual porosity will greatly influence the value of S,
and, consequently, the resulting driving pressure and permeability. This condition
can result in very complex transient effects, especially during filling and
dewatering. This problem is difficult to quantify or even model, and is an area for
future studies.

3.3.1 Preliminary pore pressure estimation

This concept will have a great influence on the investigation program and will
depend on the hydrogeological characteristics of the studied area. Moreover, the

36
type of weather is of prime importance; for example a project in a dry climate area
will present very different natural pore pressure distribution than a project in
temperate high rainfall regions.

However, unless the engineer has sufficient knowledge of the actual regional pore
pressure distribution for a particular site, it is recommended to neglect the natural
pore pressure (H0) in order to be conservative at the preliminary stage.

3.3.2 Pore pressure measurement

Natural pore pressure must be addressed during investigations with a sufficient


number of piezometers. This is done in order to establish basic parameters such as
pore pressure distribution and local pre-existing gradients, along tunnel alignment
and especially near the end of the steel liner. For instance, the result of
permeability tests, such as packer tests or Lugeon tests, depends on the initial
natural pore pressure that has to be measured when carrying the tests. This implies
that investigation works described in section 3.2.2 is interpreted in a hydro-
geological sense and that the major water bearing features (contacts, joint
families, faults, etc.) that should be encountered during tunnel excavation are
known. And that the investigations will be oriented toward finding these water
bearing structures. Boreholes can be equipped with open tube or piezometer and
sealed with a bentonitic plug to avoid surface water influence. Details of rock
piezometer installation are beyond the scope of this report, however, details can
be found in Patton (1987). These piezometers should be left operational for
continuous monitoring during the initial filling of the tunnel and to assess rock
mass saturation. Continuous monitoring will provide useful information to
validate initial assumptions, especially regarding excepted leakage.

Further measurements can be carried out while performing in-situ hydraulic


testing, measuring the natural pore pressure within the tested section prior to and
post to test (Rancourt and Mitri, 2007). This aspect will be discussed in the
following section.

37
3.4 In-situ stresses

As revealed in chapter 2, the minimum principal stress and its orientation, are the
parameters that plays a vital role in the design of unlined pressure tunnels. The
minimum principal stress is the basis of the Norwegian approach, which stated
that if the minimum stress is smaller than the water pressure around a tunnel, any
existing joints perpendicular to the minimum stress will tend to open, thus
creating hydraulic jacking conditions (Broch, 1982; Brekke and Ripley, 1987).
This assumes that there will always exist a joint perpendicular to the minimum
stress direction. Pressure tunnels are often constructed near the ground surface
where there are open and weathered joints, and where more decompressed rock
masses are encountered. For any project, the in-situ stress must be first estimated
at the preliminary stage, and later measured from the surface at the exploration
and design stage. Finally, it must be checked (if prior measurements were not
satisfactorily) during construction in order to verify final design. Rancourt and
Mitri (2007) discussed hydraulic testing methods and proposed measuring a
critical effective pressure (Pc) that would be related to the pressure at which the
flow increases non-linearly. Rather than using absolute minimum stress (σ3) in the
stress criterion for jacking (see section 2.1.1.2), the authors suggest using critical
effective pressure because, in some cases, the minimum stress given by hydraulic
jacking tests could represent the pressure that initiates hydraulic shearing. This
might occur before hydraulic jacking (see equations 2.7), and therefore lead to
false minimum stress determination.

In this report, stress anisotropy is defined as the ratio between the horizontal stress
and the vertical stress, with the following equation:

K =σ h 3.2
σv

The use of equation 3.2 is convenient for preliminary stress estimation because
the vertical stress can be assumed to be controlled by gravity, while the horizontal
would be more representative of possible tectonic stress (or the lack of it). This

38
expression also includes a shear component because σh and σv are not necessarily
in the direction of the principal stress.

Stress anisotropy has to be considered as a key parameter (Pine and Batchelor,


1984; Barton, 1986; Ming and Brown, 1988; Min et al. 2004). All these authors
address the importance of stress anisotropy on deformations and hydraulic
shearing of rock masses especially when K < 1.

3.4.1 Preliminary stress estimation

It is now common practice to perform in-situ stress measurements for any unlined
pressure tunnel project (e.g. Merritt, 1999). But at the preliminary stage, in-situ
stress measurements are not available and minimum stress and stress anisotropy
must be estimated. This estimate has to be done with regards to an extremely
daunting complex system consisting of tectonic history, jointing, topographic
variations, erosion, isostasy, weathering, etc. Careful estimations must therefore
be made using appropriate assumptions and tools. At the stage of preliminary
design, cover criteria are often used with proper factors of safety and topographic
corrections. However, the major drawback is that it may create a false sense of
security for the designer, since it is well known that the use of these criteria is
highly uncertain (e.g. Deere, 1983; Benson, 1988; Alvarez, 1997).

For a large scale start, Reinecker et al. (2005) published the latest World Stress
Map, a project that was originally presented by Zoback in 1992. The information
is applicable on a very large scale and can only provide a general idea of the
principal stress directions and magnitude in a vast region. Tectonic
manifestations, such as active faults and volcanoes, will surely have a large
impact on the local stress regime with high local stress concentrations and K
values above the unity. While old cratons, which have suffered time and glacial
erosion, are decompressed near the surface and possibly present K values smaller
than the unity.

On a smaller scale, the geological and structural features will modify the stress.
For example, soft layered sedimentary rocks will present different local stress

39
variations than hard and fractured granite. Faults (and even more local joints) will
produce high local stress variations (Martin and Chandler, 1993). Therefore, good
knowledge of the rock mass structure will be particularly important to interpret
stress measurements.

For a deep tunnel, the in-situ stress is not likely to be influenced by the surface
feature and is more controlled by tectonics, although in-situ stress variations may
still persist due to the presence of faults and dykes. However, when tunnels
approach the ground surface, the local topography becomes more important.
Figure 3.2 illustrates four typical shallow topographic conditions.

a) c)

CV

CV

b) d)

β
CV
CV*Cos β

CH
CV

Figure 3.2 - Typical shallow topographic conditions

The role of topography is well known and classic rock mechanic authors like
Talobre (1967) and Goodman (1980) gave schematic examples of stress variations
around valleys. The role of topography was studied by Ripley and Brawner (1990)
who, through the use of case histories, discovered that steep terrain minimum

40
stress can actually be much less than predicted by gravity method. Amadei and
Pan (1995) studied stress variations induced by topography using analytical
solutions and concluded that topographic features create low stress (even tensile)
zones which have a large influence on unlined tunnel behaviour. Chapter 4 will
present numerical modelling results for each of these topographic conditions.

Flat

As revealed by Brady and Brown (1993), stress to vertical cover relation near the
surface is rarely satisfied and the vertical direction is rarely a principal stress
direction. Hoek and Brown (1980) and later Arjang (1989), presented the
following relations between measured in-situ stress and depth:

σ v = 0.027 z (3.3)

100
K min = + 0.3 (3.4)
z
1500
K max = + 0.5 (3.5)
z

These relations illustrate that vertical stress seems well correlated with depth z
(in m), but is not the case for horizontal stress, where at low depth, K values range
from 0.5 up to 3.0. However, these relations are given for depths as low as 3,000
m, for shallow structures (< 200 m), the results present a very high scatter. Noting
that the aforementioned equations are derived from measurements made largely
for underground mines, there is a specific need for future research to be focused
on studying local in-situ stress variations at shallow depth.

Slope

The specific conditions of rock masses in the vicinity of slopes was addressed by
Terzaghi (1963) and are stated to be dependent on the type of rock and the
geological history of the slope. Exact analytical stress distributions for finite
elastic slopes are given by Silvestri and Tabib (1983). They provide influence
diagrams of stress according to slope angle and slope height. Ripley and Brawner

41
(1990) made a comparison between estimated vertical stress and actual measured
stress for several available case histories with steep topography.

They revealed that for steep topography, vertical rock cover provides an over-
estimated value that can lead to serious consequences. This is consistent with
findings from several authors who have found much lower minimal stress than
previously expected (Bouvard and Pinto, 1969; Deere, 1983; Marulanda and
Gutierrez, 1999; Dalho et al. 2003; Rancourt et al., 2006a). This result also
validates the slope correction initially proposed by Berg-Christensen and
Dannevig (1971) and is presented in table 2.1.

Hill

The decompressed nature of hill and protruding noses was acknowledged by


Broch (1984) and reiterated in Brekke and Ripley (1987). These topographic
features are generally made of harder rock masses (compared to the valley
bottom) with strong horizontal stress relaxation. Consequently, contours between
depressions should be trimmed when assessing minimum rock cover (see
table 2.1).

Valley

According to several authors such as Talobre (1967), Goodman (1980), Brekke


and Ripley (1987), Hartmair et al. (1998), the presence of a deep valley has a vast
influence on local and near surface stress. James (1991) studied the stress changes
caused by valley erosion. The author examined the progression of river erosion,
leading to continuous slope failures and equilibration of gravity and tectonic
stress.

Valley bottoms are likely to correspond with softer erodible geological features
compared to the surrounding hard mountain tops. At the bottom of valleys, rock
has almost no vertical stress but has a large horizontal component due to the
presence of high valley walls on both sides. Valley rock is squeezed horizontally
like an abutment. At mid-elevation of valley walls, principal stress is parallel to

42
the ground surface and at the top of valley sides, the surface is decompressed and
has low horizontal stress.

3.4.2 Stress measurements

In-situ stress measurements are now routinely performed to verify the


confinement criteria assumed at the preliminary stage and during early design and
also in order to assess stress heterogeneity and stress anisotropy as much as
possible. However, this kind of in-place testing is always tainted by high
variability due to field and interpretation inaccuracies. Consequently, results have
to be regarded as a range rather than as absolute values. Several methods exist to
measure the in-situ stress in rock masses; the present thesis will consider only two
of them: stress relief methods and water injection methods. It is always desirable
to carry both methods for the same project since this allows for the comparison
and validation of results for the final design (Rancourt et al., 2003).

3.4.2.1 Stress relief methods

Stress relief methods are based on elastic theory and use strain gages to measure
deformation while rock is relieved from natural stress with a drill or a saw. Stress
is back calculated using deformation parameters of the intact rock sample; this
allows measuring the total stress in intact rocks.

This method may provide the complete 3D principal stress tensor (e.g. Leeman,
1969, Gill et al., 1987). These methods involve small volumes of intact rock and a
model was developed for anisotropic non linear elastic rocks (Corthesy et al.,
1993). If the rock mass is highly fractured, measurements are difficult to obtain.
This kind of method requires intact rock borehole sections and corrections needs
to be carried out to account for borehole stress modification effects (Rahn, 1984).

This type of method is a total stress method usually done deep within rock masses
where natural pore pressure is negligible. However, existing pore pressure might
be drained in the vicinity of boreholes, especially when measures are performed

43
from a tunnel, whereby the stress results would represent the effective stresses
(Dalho et al. 2003).

3.4.2.2 Water injection methods

Water injection methods are very useful for pressure tunnel design because they
recreate, at borehole scale, the physical process happening in pressure tunnels
(Doe and Korbin, 1987). They can provide two fundamental parameters: in-situ
rock mass permeability and in-situ natural confinement. Permeability and
effective confinement can be estimated at different scales in the rock mass, using
water pressure injected in a borehole section at different lengths.

For permeability, several methods are available and are based on the early work
done by Maurice Lugeon, who established the principles of the water pressure test
in a section of borehole, with the objective of determining rock mass groutability.
Later, Louis (1974) proposed a method to measure permeability using two
boreholes, one injection point and one observation point. This principle was
generalized by Hsieh and Neuman (1985a, 1985b) for the determination of the
three-dimensional permeability tensor for an anisotropic media using borehole
field information called the cross-hole testing.

For in-situ confinement, two types of tests were developed: the hydraulic
fracturing (HF) and the hydraulic testing of pre-existing fractures (HTPF). The
first methods (Haimson and Fairhurst, 1967) were developed from deep vertical
boreholes for the petroleum industry. They had to make the assumption that the
borehole was parallel to one of the principal stresses. Later, Cornet and Valette
(1984) introduced the HTPF method to calculate the complete 3D stress tensor
using an arbitrary oriented borehole and at least 6 differently oriented test
fractures.

Dalho et al. (2003), revealed the great influence of the existing drainage situation
on the vicinity of test sites on stress estimation results. All these effects are
included in the raw test results and must be carefully interpreted in order to
provide a clear picture of the permeability and in-situ stress regime. Research is

44
still needed in this domain in order to obtain a more comprehensive interpretation
of test results and possible negative external influence.

For all water injection methods, the use of results must be interpreted with care
and there are major drawbacks which are (Louis, 1974):

1-Effect of the radial flow (variation of the velocity in the flow direction)
2-Turbulence effects
3-Deformation of the medium under joint water pressure during the test
4-Influence of the permeability of other joints set on the tested one
5-Head losses
6-Influence of time and possible unsaturated zones

Both the HF and the HTPF methods were reviews by Haimson and Cornet (2003)
in the Rock Stress Estimation Special Issue.

Hydraulic fracturing

For hydraulic fracturing the test section is chosen free of existing rock joint and
the section is pressurized until a fracture occurs and the subsequent pressure
required to hold the fracture open is called the shut-in pressure. It is assumed that
the new fracture orientation is perpendicular to the principal minimum stress.

Figure 3.3a shows a typical shut-in curve and Ps1 is the shut-in pressure point as
defined by the inflection point method (Gronseth and Kry, 1981) and located by
the tangent divergence. As for Ps2, it represents the tangent intersection point and
usually it is a lower bound for shut-in point.

There are several methods that can be used to interpret the shut-in pressure and
this was studied by Aggson and Kim (1987), Haimson (1993) and Guo and al.
(1993). All authors agreed that the inflection point method initially used by
Haimson and Fairhurst (1967), always produces high shut-in pressure, especially
for low stress. Since this method does not give a conservative estimate of

45
minimum principal stress, the intersection of the tangents would provide a more
conservative estimate.

Hydraulic jacking

For hydraulic jacking, a test section is chosen with one or more existing joints and
the section is subjected to incremental pressure steps; usually the test pressure
must exceed the future internal tunnel driving pressure. Hydraulic jacking is a
variant of the standard Lugeon pressure test where the pressure range is
sufficiently large to open fractures. It is also regarded as a direct test of the
confinement capacity of the rock mass to withstand internal pressures (Broch et
al. 1997). While performing these tests, it is important to make sure that all
possible joints are tested with different orientation. It would also be of great
interest to measure the tested joint orientation. Figure 3.3b illustrates a typical
hydraulic jacking curve (e.g. Doe and Korbin, 1987). The curve is divided into
three parts by critical effective pressures Pj1 and Pj2. The first part, when the
pressure is lower than Pj1, is a linear pressure/flow relation appearing at low
pressure and controlled by initial hydraulic aperture and contact areas of the
fractures (Rutqvist and Stephansson, 1996). The slope (k1) for this part of the
curve is proportional to the permeability of the tested section. Pressure Pj1
represents the minimum jacking where the curve becomes non-linear. Between
Pj1 and Pj2 the slope varies and is controlled by the normal fractures effective
stiffness and their normal effective stresses. Effective pressure Pj2 is the ultimate
jacking; beyond that, fractures are jacked open and the global stiffness equals the
surrounding rock mass stiffness (k2).

It was observed that fractures will begin to open for an excess of fluid pressure
even if the current fluid effective pressure is less than the total in-situ stress
normal to the fracture plane (Rutqvist and Stephansson, 1996). This behaviour
was also observed experimentally by Cornet et al. (2003) and Bart et al. (2004)
and suggest possible influences on the resulting test curve. Test equipment is
basically the same as the hydraulic fracturing test, which suggests that effective

46
pressure Pj1 is a conservative value to be used. And it is equal to the critical
effective pressure Pc and is the minimum stress used in the design.

Several authors agreed on time effects on the response of water pressure tests due
to the variation in the tested fracture geometry (area of contact, saturation, etc.),
therefore showing a large hysteresis along with testing time (e.g. Cappa et al.,
2005; Bouvard and Pinto, 1969).

Pressure (P) Flow (Q)

Linear P vs Q
k2
1
Instantaneous proportional to fracture
Ps1 Shut-in hydraulic aperture
1 2 3
2
Non-linear pressure
drop in relation with the the effective normal stress
fracture elasticity and the
effective stress Hydraulic jacking zone
3
proportional to surrounding
rock mass elasticity and
Ps2 k1
effective stress

Time (t) Pj1 Pj2 Pressure (P)

a) TYPICAL SHUT-IN CURVE b) TYPICAL JACKING CURVE

Figure 3.3 - Typical hydraulic testing curve (from Rancourt and Mitri, 2007)

Influence of permeability

As previously discussed, the shape of a test curve could depend on rock mass
elasticity as well in-situ stresses. For a soft and highly deformable rock mass,
hydraulic jacking might occur gradually during the elastic deformation of the
fracture. On the other hand, for a stiff hard rock mass, where strains are small and
stress is high, jacking occurs as an irreversible and sudden fracture failure that
increases the opening and leakage. Both situations result in the same excessive
leakage. Figure 3.4 illustrates the typical effect of initial fracture permeability on
hydraulic jacking curves.

47
Performing a test in a high permeability fracture provides smoother curves with a
less sharply define non-linearity and might lead to subjective jacking pressure
determination. If a test is too pervious, there is not enough pressure to initiate
hydraulic jacking and the test fails to provide critical pressure (Pj1). The test is,
however, very informative about the permeability of this borehole section.

Marulanda et al. (1990) observed that when testing highly permeable joints, the
pump used needs to be powerful enough in order to pressurize the zone and obtain
a reliable jacking curve. In this case, the highly permeable zones were primarily
grouted in order to later be able to build up pressure and measure the hydraulic
jacking value.

Flow (Q)

y
ilit
e ab
p erm
re
ctu
fra
gh
Hi
ity
Low fracture permeabil

Pressure (P)

Figure 3.4 - Typical tendency for permeable and impervious rock masses
(modified from Marulanda et al., 1990)

Influence of initial pore pressure

As discussed in section 3.2, knowledge of the naturally existing pore pressure


around the tunnel alignment is to be considered. Hydraulic testing results are
influenced by initial pore pressure in the tested area. As previously stated, the
conventional Lugeon test requires knowledge of the pore pressure to be properly

48
interpreted. Cornet and Valette (1984) mentioned the possible influence of natural
pore pressure on HTPF test results. Conversely, HF test interpretation assumes the
tested zone to be dry and not influenced by in-situ pore pressure. Accordingly,
what is measured during hydraulic testing is the total stress, which includes the
effective stress acting perpendicular to the fracture and the natural pore pressure.

Recent work on this subject was presented by Dalho et al. (2003), with results
indicating that minimum stress was reduced by drawdown in pore pressure
following tunnel excavation. Assuming that since stress and pore pressure are
coupled according to the Biot’s poro-elasticity theory (e.g. Wang, 2000), because
of the high stiffness of the rock mass, the drop in pore pressure will generate a
tensile stress component that will disturb initial stress. In light of this, the testing
methodology should be adapted accordingly. Figure 3.5 reveals two types of
hydraulic testing set-ups: one from the surface before tunnel excavation and the
other from tunnel level after tunnel excavation.

Testing from the surface generally requires deeper holes with more costly testing
due to the depth of the tested zone. However, it would not be influenced by the
drawdown in pore pressure and the possible stress alteration. On the other hand,
underground tests are convenient due to short holes and the possibility to drill
several holes at low cost. For this type of set-up, the tested section cannot be
located exactly at the tunnel location, a standard practice in surface tests. Test
sections also have to be placed far enough from the excavation to be separated
from the deformations induced by the tunnel. Furthermore, as previously
mentioned, the tests might be influenced by pore-pressure drawdown of the
surrounding tunnel.

49
a)

Assumed regional
groundwater level

Tested zone

Futur pressure tunnel

b)

Assumed regional
groundwater level

Existing pressure tunnel


Tested zone

Figure 3.5 - Examples of testing situations, a) testing carried from the surface
before tunnel excavation, undrained, b) testing from inside
the tunnel with a drained rock mass

3.5 Rock mass permeability

For a pressure tunnel it is of great interest to estimate the total expected leakage
and also to locate the risk of occurrence of local water outflow. Local water
outflows are usually associated with structures like plugs and shafts, and may be
primarily controlled by open joints or local high permeability zones due to high
rock mass deformations, while the total average water outflow may be the sum of
several zones, with an equivalent and average permeability over each zone
Rancourt et al. (2006b). This method assumes that the rock mass behaves like a
Darcy-type medium and the permeability needs to be preliminarily estimated.
Assuming permeability and natural existing pore pressure values along the tunnel,
it is then possible to estimate the expected total leakage in or out of the tunnel.

50
As presented in chapter 2, water leakage is one of the two most fundamental
concerns, the other being tunnel stability. With high leakage levels there can be
considerable loss of revenue, as well as leakage toward the powerhouse; leakage
reaching the surface could induce landslides (soils or rocks), etc. According to
Jiao (1997), the principal parameters influencing leakage are in-situ stresses,
effective water pressure and joint aperture (permeability).

The distribution of existing natural water pressure around the tunnel is of


particular importance, especially near the steel liner, the plugs and the surge shaft.
However, for jointed rock masses, difficulties arise due to the presence of local
high permeability zones and faults and open joints, which behave differently than
conventional porous materials. As a result, the engineer must understand both
total water outflow and pressure distribution around the tunnel in order to make a
sound design.

Fluid flow and pressure distribution in rock masses are generally defined in one of
three distinct concepts:

1- Open joint, fault and shear zone of large scale


2- Equivalent continuum system
3- Porous intact rock matrix

These concepts are based on the existence of three different porosity types. Junriu
et al, (2004) even proposed a quadruple porosity model, where concept no. 2 is
divided in two parts. The form of the model largely depends on the scale of the
examined area, if one is studying a local structure with specific geological
features such as discrete open joints or permeable faults. Concept (1) therefore
controls the pressure and flow. In the case of a whole tunnel evaluation, which
represents a large scale structure, the rock mass may likely behave like an
equivalent fractured rock mass continuum (2). Except for porous sandstone and
several other rock types, concept (3), rock matrix porosity, can generally be
neglected for hard rock because of the small contribution to global mechanisms.

51
In addition, because at the scale of tunnel projects, rock mass is always jointed to
some extent (Louis, 1974).

3.5.1 Open joints and faults

Open joints or faults or other highly permeable large geological structures are
controlled by the cubic law and the principle of water balance, as defined in
equation 2.4. This concept was not studied in this thesis.

3.5.2 Equivalent fractured continuum

Because pressure tunnels are generally large structures it can be assumed that the
rock mass behaves like an equivalent fractured media (see section 2.2.1). In this
thesis, leakage is associated with permeability through the Darcy law, where the
basic permeability versus flow relation is given in equation 2.1. This equation is
used throughout this study, with different concepts that apply to the study of water
flow through fractured rock masses. However, the continuity assumption needs to
be verified on the field with the aim to locate local permeable structures.

3.5.3 Preliminary permeability estimation

The estimation of preliminary permeability is a difficult task that can only be


realized using existing historical data or the rock type from regional mapping and
linking it to empirical values for similar rocks. Structural geology can also be
used to identify possible high permeability zones. If the rock is anisotropic, there
is a good likelihood that the rock mass will present an anisotropic permeability
tensor. Figure 3.6 illustrates the typical permeability according to rock and soil
types.

52
Figure 3.6 – Typical permeability ranges for rocks and soils
(Freeze and Cherry, 1979)

3.5.4 Permeability measurements

3.5.4.1 Equivalent continuum

The topic of water injection was already addressed in section 3.4.1.2 and this
section will only present specific aspects pertaining to pressure tests and its use in
estimating permeability. Pressure tests are also known as packer tests or Lugeon
tests (Lugeon, 1933). The method to measure equivalent continuous permeability
is to carry out pressure tests along an isolated borehole section. Tests are usually
done in stepped effective pressure while reading the water flow. Each pressure
step is allowed to stabilize for 5 to 10 minutes, after which the flow is noted for

53
the particular pressure. Originally, a Lugeon unit was defined as the amount flow
(in litres per minute) under a pressure of 1,000 kPa per metre length of the tested
zone.

Results are plotted on a pressure versus flow graph similar to the one in
figure 3.6b. The results have to be interpreted in order to detect any particular
behaviour that would possibly influence the results. The graph provides
information on the hydraulic properties of the rock mass. Figure 3.7 illustrates
typical curve types with specific interpretations from Widmann (1996).

Q (a) Q (b) Q (c)

P P P

Q (d) Q (e) Q (f)

P P P

Figure 3.7 – Typical relations of effective pressure against water flow


(modified from Widmann, 1996)

54
Each curve type represents characteristics of a specific behaviour:

(a) laminar flow


(b) turbulent flow
(c) hydraulic fracturing or hydraulic jacking
(d) erosion of joints
(e) plugging of joints
(f) probable flow around the packers

The estimate of k is reliable for linear curves (a), but for other curve types, care
must be used in the interpretation of tests. When the curve is linear, the Lugeon
value is easily obtained and is proportional to the slope of the curve, but when the
curve presents non-linear aspects such as for (b), (c), (d), (e) and (f), an
interpretation must be made in order to obtain a permeability value. Ewert (1994)
presents a discussion regarding curve interpretation and concludes that some
results can be quite ambiguous because a variety of geological settings can yield
similar results. He recommends sticking with graphs that are clearly different and
which can be geologically explained. In addition, water flow is not necessarily
proportional to the number of tested joints, or the rock mass quality; one open
joint in hard granite might take more water than several tightly closed joints in
shale (Foyo et al., 2005).

Test results are highly influenced by borehole orientation, according to specific


pervious joints. Consequently, borehole orientation should be optimized in order
to intersect the joints that are expected to be pervious. For an equivalent media,
the tested length should be adjusted to include several joint families.

Because of these aforementioned reasons, the results are not likely to give an
absolute permeability value but rather a relative index permeability value.

Since the early work of Lugeon several expressions were present in the literature
(e.g. Hvorslev, 1951; Ballivy and Bouja, 1992,), and Singhal and Gupta (1999)
have proposed the following simplified equation:

55
Qc / LB
kc = 1.12 (3.3)
H

It is assumed however that the extent of the cone of influence (formed by the rise
of the water table during the test) is very large compared to the borehole diameter.
The test should be carried in saturated zone, the pressure should be kept constant
until the flow stabilise.

Nonetheless, the conversion to permeability is based on the spherical flow


concept developed for soils. This concept is not easily transferable to jointed rock
masses with several numbers of joints.

Finally, the types of curves obtained in figure 3.7 are highly related to the
maximum effective pressure of the tests. Ewert (1994) recommends that the
maximum effective pressure be at least 20% higher than the working pressure of
the structures. Furthermore, some tests should push the pressure high enough in
order to examine the critical pressure defined in section 3.3.2.2, which initiates a
change of slope in the curve, for the purpose of testing available rock mass
confinement. As a result, the pressure test would combine relative permeability
and critical effective pressure into one hydraulic test (Rancourt and Mitri, 2007).

3.6 Slope stability

Slope stability is a major aspect of tunnel stability and requires a comprehensive


structural analysis based on joint mapping and stereonets for basic pressure tunnel
design. The preliminary identification of major structural features is also
necessary for tunnel alignment and investigation purposes. These structures are
commonly faults, shear zones, geological contacts, dykes, beddings and joint
families.

Tables 3.5 and 3.6 present some typical cases that have to be addressed in the
design process, by first identifying if the case is possible, and then by using the
design methods discussed by Ming and Brown (1988) and by Alvarez (1997)

56
amongst others, to evaluate rock mass stability. The methods are briefly described
in the figures below.

Table 3.5 - Failure by sliding along slopes

Schematic representation General Mechanisms References


- Possible slide of a Rock Slope Stability
block wedge Terzaghi (1963);
- Water pressure Hoek and Bray (1974);
destabilizing Pressure Tunnels
- See Illinois Univ. Alvarez (1997)
method in table 2.1

- Possible landslide General Overburden


- Water pressure Slope Stability
destabilizing e.g. Abramson et al.
(2002)

57
Table 3.6 - Failure along fractures

Schematic representation General Mechanisms References


Flat topography
a) b) a) Jacking with low Ming and Brown
horizontal stress (1988);
(K<1) and/or high Alvarez (1997)
vertical shear stress
b) Pervious vertical
structure bringing
pressurized water in
horizontal near
surface structure
(Critical for K>1)
Slope topography
a) b) a) Jacking with low Ming and Brown
parallel to ground (1988);
surface stress Alvarez (1997)
b) Pervious structure
bringing pressurized
water in parallel to
surface structure,
possible buckling of
rock slab
Valley topography
a) b) a) Usually not James(1991);
problematic because Hartmaier et al.
of high horizontal (1998)
stress at the valley
bottom
b) Jacking or buckling
of horizontal valley
bottom slab

58
3.6.1 Preliminary structural assessment

Once again, in the early stages of a project, very little information is available and
only broad regional mapping can be used. As a result, a preliminary structural
assessment must be done based on regional geology, regional hydrogeology and
knowledge of structural features. At this stage, however, contingency risks are
high and provisions should be made for unforeseen and unfavourable structural
aspects that are likely to be found during site investigation.

3.6.2 Detailed structural investigation

Structural investigation is usually carried out first by a good structural mapping of


existing outcrops on the project site and by several surface drill holes. With this
field information, stereonets can be prepared and major joint families can be
identified. This work allows the preliminary study of possible wedge stability.
Also at this stage, site partitioning can be initiated for subsequent investigations,
which can include geophysical methods, but will surely include several other
boreholes allowing to obtain data at tunnel level. Borehole logging should include
fracture orientation or a geocamera to extract the most useful information
possible.

3.7 Basic parameter interactions

The interactions between parameters within a physical system is called coupling,


and it is increasingly being considered in the evaluation of pressure tunnel
behaviour (Alvarez, 1997). To achieve the coupling of parameters, one has to
account for the basic chain of interactions; for example ΔP → Δe → Δk → ΔQ,
whereby a change in water pressure leads to a change in joint hydraulic aperture
which, consequently, changes the rock mass permeability and then the total water
flow. Schleiss (1986) was one of the first to address the influence of fracture
deformation on rock mass permeability due to internal water pressure tunnels. For
pressure tunnels, apart from the first mechanical deformations of tunnel openings
following excavation, coupled interactions are basically of two types: drainage
effect of the empty tunnel (construction, dewatering) and tunnel pressurization

59
(filling-up, saturation). The analysis becomes more complex if one considers the
fact that hysteresis effects were observed when comparing the loading and
unloading of joints (Kranz et al., 1979; Bart et al. 2004).

The present thesis attempts to investigate the concept of permeability variation


according to water pressure and its effects on the total leakage around a tunnel.
More specifically, it aims to study when the coupling of permeability and water
pressure is significant and should be taken into account during pressure tunnel
design. Several models exist to address the coupled interaction between
permeability and/or porosity and volumetric strain and/or displacements. The
following table provides the coupled relations.

60
Table 3.7 - Summary of coupled permeability models

Model Controlling Relation Model type Limitations


Parameter
Schleiss Radial ⎡ ⎤ Radial, -Valid for high fracture
(1986) displacement ⎢ ⎥ Isotropic density
⎢ γ wκ 3 v ( r ) 3 ⎥
kc = k0 + ⎢ 2 ⎥
, -Requires radial
⎢ 3⎛ 1 1 ⎞ ⎥ displacement around the
⎢ 6 μr ⎜⎜ s + s ⎟⎟ ⎥
⎣ ⎝ 1 2 ⎠ ⎦ tunnel

⎛ E rm ⎞
κ = 1 − ⎜⎜ ⎟⎟
⎝ Eir ⎠
Lombardi Volumetric m Isotropic -For fissured elastic,
⎛ Δε ⎞
(1989a) strain k c = k 0 ⎜⎜1 − ⎟⎟ saturated rock masses,
⎝ ε0 ⎠
subject to compressive
stress and low shear
stress.
Chen and Normal and (e0 + Δv )3 Anisotropic, -Consider joint family
Bai (1998) shear strain k c = 12si Discrete separately
(for a joint set) fissures -Must know joint family
spacing
-Not dependent of k0
Liu et al. Normal and 3 Anisotropic, -Consider joint family
⎡ 2(1 − Rm) ⎤
(1999) shear strain k c = k 0 ⎢1 + Δε ⎥ Discrete separately
⎣ n ⎦
(for a joint set) fissures -Must know and assume
porosity constant

As illustrated in table 3.7, all models are more or less based, or at least inspired
by, the cubic law (eq. 2.2). The Schleiss (1986) model is a radial model and
requires radial displacement around the tunnel. Lombardi (1989a) and Liu et al.
(1999) introduced models which use volumetric strain as the main parameter. The
model by Chen and Bai (1998) introduced displacement perpendicular to fracture
sets.

61
CHAPTER 4

NUMERICAL MODELLING OF SHALLOW UNLINED PRESSURE


TUNNEL

In this chapter, numerical modelling is carried out to investigate principally the


preliminary design assumption that is made using gravity criteria. Hence, the
influence of topography, the rock cover versus the tunnel diameter and the
influence of structural features were modelled and checked for robustness of
existing gravity criteria. Also some preliminary modelling was done to study the
rock mass deformability and rock mass permeability coupling significance on the
preliminary design of unlined pressure tunnel. Criteria used in the parametric
study are presented in this chapter. Table 4.1 presents a summary of numerical
modelling done using equivalent continuous rock masses and performed using
basic generic cases. Appendices A, B and C are presenting the models parameters,
numerical codes and basic geometries respectively.

Table 4.1 - Summary of numerical modelling studies

ROCK MASS MODELS OBJECTIVES


Elastic with effective cover 1.1 Basic tunnel geometry Validate the use of circular tunnel geometry
criteria assumption 1.2 Rock cover versus tunnel Propose limits for C/Dt
diameter
1.3 Types of topography Validate criteria
A. Flat Validate criteria
B. Slope Propose correction angle limit
C. Hill Criterion validation
D. Valley

1.4 Rock mass anisotropy Study anisotropy effect on effective cover


1.5 Structural features Study structural feature effect on effective
cover
1.6 Comparison models Show the differences between models with
C. Hill and without structural features. And
D. Valley investigate the presence of overburden in
the valley
1.7 Rock mass plastic behaviour Compare elastic and plastic models in terms
Plastic with effective cover of effective cover criteria assumptions
criteria assumption

Elastic with hydro-mechanical 1.8 Influence of hydro- Flow study and leakage assessment to
interactions mechanical coupling propose modulus value where coupling
becomes significant

62
4.1 Numerical modelling method

A wide range of commercial numerical codes are available for the study of stress
and fluid flow for discrete rock joints as well as for equivalent continuum rock
mass. Programs can perform two-dimensional (2D) and three-dimensional (3D)
analysis of stress, displacements and water flow.

The 2D models are used generally for linear excavation, such as tunnels and
shafts, assuming that the stress and displacement in the plane normal to the axis
are not influenced by the adjacent excavation. In the case of short tunnel or more
complex underground structures like powerhouses, bifurcations and other three-
dimensional openings, the use of 3D models is required.

Finite Element Method (FEM)

Finite Element Method is the basic method of defining a domain with boundaries
and dividing the domain into a finite element mesh. The domain can be
heterogeneous and/or anisotropic. Problems involving non-linear properties can
also be solved. The elasticity module to be introduced in the model to describe the
rock mass is strongly influenced by the scale of the problem. This method was
used by Brekke et al. (1970) to study pressure shaft failure.

It is a well known method with powerful commercial codes (COSMOS,


ABAQUS, etc.). These programs allow all types of loads, sequences of
construction, support installations, etc. Numerical divergence usually results in
the model predicting obvious anomalous physical behaviour. However, because, it
requires considerable effort in data preparation for a problem and might require
costly computing, it may not be well suited for modelling infinite boundaries.

Finite Difference Method (FDM)

This method is another domain technique that requires the rock mass to be a
continuum and that the problem be stated as a differential equation. This method
assumes that for a small time step, disturbance at nodes within the rock mass

63
influence only its immediate neighbours. Node displacements propagate through
the model over several time steps until convergence is reached.

These programs allow time dependant analysis where the user can follow the
evolution of the system. It is also appropriate to model large deformations. FLAC
is a well known commercial code using this method.

Boundary Element Method (BEM)

This is a boundary method where only the boundary of the excavation is to be


analysed and needs to be defined. External boundaries are not defined and rock
mass is represented as an infinite continuum. The excavation boundary is divided
into elements and the surrounding rock mass represented implicitly. Stress inside
the rock mass is extrapolated from the boundary solution.

Unlike the FEM, the matrices of equations used in this method are not banded and
symmetric. This method does not require a whole mesh and only the boundary
elements are needed, thus creating fewer elements than FEM or FDM methods.
Joints can be modelled using the displacement discontinuity approach; however, if
all joints are explicit models, an increase in computational effort is obtained.

Discrete Element Method (DEM)

This method regards the rock mass as an assembly of blocks that, if loaded, will
deform, rotate and translate. The rock mass is explicitly divided into zones with
the deformations governed by the contact between blocks. The elasticity of the
block itself is generally neglected but some codes allow it to be considered.
Contacts may be represented by the overlaps of adjacent blocks and large
displacements may occur.

This method is well suited for modelling blocky rock masses with structurally
controlled types of failure. This method requires high user expertise and
significant computational time and effort, especially for complex models. Finally,

64
it is very difficult to estimate representative input parameters. UDEC is a known
commercial code which uses this method.

4.2 Selection of modelling tool

In the selection of a modelling tool, the user must choose amongst a large variety
of codes depending on the type of problem to be solved. Other considerations,
such as cost and availability, might sometimes have an influence on this choice.

For unlined pressure tunnel design, the Norwegians were the first to use a finite
element method to verify the slope correction of the criterion (e.g. Brekke and
Ripley, 1987). More recently, new codes became available, taking water pressure
effects into account, and Alvarez (1997) used the hydro-mechanical capacity of
the UDEC DEM code to numerically model both structural and hydraulic aspects
of pressure tunnels. The principal result from Alvarez (1997) was presented in
section 2.3 and is a criterion that considers structural features near a slope. The
criterion is used in the design table presented in chapter 6.

Considering the work already done with a discrete code, it was then decided to
use a continuous code with hydraulic modelling capacities. The study would also
be 2D for the sake of simplicity and would concentrate on shallow tunnel
situations with typical topographic influences.

The code to be used should allow the user to specify the loading conditions, the
geometry, the elasticity and the permeability of a rock mass, and then solve the
model to evaluate the stress, strain, displacement, pore pressure and flow inside
the model. Keep in mind, however, that programming capacities are required to
model coupling effects.

The basic loading approach is the following: for a horizontal topographic surface,
the rock mass is loaded with the assumed in-situ stress, as well as the overburden
(σv) and tectonic forces (σh). In the case of inclined slope, hill or valley
topography, the model should attempt to include the effect of erosion, and its
unloading effect near the surface. Further to initial in-situ loading and subsequent

65
erosion, the excavation is then made and the loads are redistributed around the
zone of influence of the excavation.

For stress analysis, the deformations are assumed to be elastic and are basically
related to the rock mass quality. In the case of flow analysis, results are influenced
by rock mass permeability and the study includes the analysis of the interactions
between deformations and permeability.

In order to be able to carry out this type of analysis, the FLAC modeling code was
therefore selected for the model parametric study.

4.2.1 FLAC 2D

FLAC2D Version 4.0 is a general purpose continuum program for the modeling
of problems in rock and soil mechanics. The program uses a hydro-mechanical
interaction module, and has an internal compiled language (FISH) which allows
users to modify applications and graphics, as well as write their own material
models. It was developed by the ITASCA Consulting Group in Minneapolis. It
has the ability to perform irregular external boundary shapes to represent
topographic variations. Hakami (2001) used FLAC3D to model the pumping test
at Sellafield and to predict hydro-mechanical consequences of shaft sinking. The
real power of FLAC is that the user can follow the process one step at a time and
view the results as the system evaluates. Numerical codes are listed in
appendix B.

4.3 Characterization of input data for pressure tunnel models

Numerical modelling is carried out to investigate important aspects of pressure


tunnel design, the verification of the reliability of cover criteria with the influence
of topography, the tunnel diameter and the structural features on preliminary
tunnel layout; rock mass deformability, rock mass permeability coupling
influence and significance on the preliminary prediction of unlined pressure
tunnel performance.

66
As for the influence of topography and stress analysis, the modelling results were
systematically compared with the gravity criteria of section 2.3. Results from
these models will be used for the preliminary tunnel layout assessment. As for the
prediction of tunnel leakage, relative flow results were compared together when
changing the deformability characteristics of the rock mass.

The important parameters are discussed in this section. Some modelling


developments are also presented for each characteristic parameter.

4.3.1 Model geometry

4.3.1.1 Tunnel geometry

Tunnel geometry varies depending on operational requirements and construction


equipment. The maximum tunnel width is largely influenced by rock mass
quality. In figure 4.1, four tunnel sections are presented: section A circular,
section B elliptic with a vertical major axis; truncated to provide a floor, section C
is close to a vertical rectangle, and section D is usually limited to bifurcation.

a) b) c) d)

Figure 4.1 – Typical tunnel geometries

Sections B and C are definitely the most used for hard rock pressure tunnels with
the drill and blast method. However, in the present preliminary numerical study,
section A is preferred because it has the advantages of symmetry and simplicity.
Verification was carried out to investigate if there were significant differences in
the model results between sections a) and c). The two tunnel geometries were
modelled and compared, when varying the ratio C/DT from 1.5 to 50. Chapter 5

67
will reveal that the differences between models with section a) and models with
section c) are negligible.

Figure 4.2 presents the details of the geometric assumptions used in the model
verifications. Although pressure tunnel width usually is less than 15 m, width (or
diameter) was varied from 2 m to 25 m in the models, in order to obtain a large
range of C/DT values.

0.25 W
r

1.4 W

0.75 W

W r = 0.65 W
D T

Figure 4.2 – Tunnel geometries

4.3.1.2 Tunnel diameter versus rock cover

The tunnel diameter relation with the height of the rock cover is not addressed by
conventional gravity criteria. This aspect was modelled with the objective of
obtaining a minimum value range for the ratio C/DT (rock cover/tunnel diameter)
to be used in preliminary design. The ratio C/DT was varied from 1.5 to 50 (1.5, 2,
2.8, 3.6, 4.5, 6.6, 9.5, 12, 16, 24, 50) and the stress ratio (horizontal to vertical in-
situ stress) fixed at K = 1. Figure 4.3 shows the models schematic layouts.

68
Cv Cv

DT
r

Figure 4.3 - Models schematic layouts

4.3.1.3 Topography

A preliminary analysis was carried out with four typical types of topographies: the
flat lying ground, the inclined slope, the hill and the valley (figure 4.4). Existing
design criteria for these types of topographies are summarized in chapter 2 (table
2.1). The Norwegian approach has long introduced the correction for slope with
the cosine reduction (see table 2.1). However, the topographic shaving method for
hill sections (Broch, 1982), does not address the shape of the hill. This aspect was
modelled to investigate the hill slope (figure 4.4). This hill shape slope should
have an influence on the depth of the correction. Recommendations regarding the
correction are presented based on model results. Note that no specific
recommendations exist for valley crossing situations.

Where the study on the influence of topography and of stress is concerned, four
types of topographies are used to assess the influence of the ratio C/DT (rock
cover/tunnel diameter), as well as the slope for hills and valleys (and how it
affects the shaving correction in the case of hills).

69
On the other hand, the tunnel leakage study only uses flat ground geometry to
study the influence of rock mass permeability and rock mass elasticity on tunnel
leakage.

a) c)

CV
y
x
CV

b) d)
y
β x

β
CV
CV*Cos β

CH
CV

Figure 4.4 – Type of topography modelled

The shape of the hill/valley is based arbitrarily on the following equation:

8a3
y= 4.1
x 2 + 4a 2

Table 4.2 presents the values of parameters a used to control the geometry of the
hill/valley models. Figure 4.4 shows the location of the coordinate systems and
figure 4.5 shows the different geometry used in the models. Parameter a controls
the shape and the slope.

70
Table 4.2 - Parameters controlling the geometry of the hill/valley models
a β (°) Remarks
15 64 Steep hill
25 52
35 42
50 32
70 25
100 17 Gentle hill

45o

C 5o

30o

DT
DT

17°
25°
32°
42°
52° 64°

52°
64°
C 42°
32°
25°
17°

Ccorr
C

DT DT

Figure 4.5 – Schematic geometry of models with C/DT ratio


and topographic variations

71
4.3.2 Structural features

Structural geological features were also introduced in models to study their effects
on the stress regime around tunnels and the variation of minimum stress. These
features are soft rock zones (low E2) and hard rock zones (high E2), and the
geological contacts (E1>>E2). Models were made using a constant C/DT ratio
of 5.

Preliminary runs were carried out using vertical and horizontal soft rock zones.
On the first series of models, the distance of the soft zones was varied with a fixed
zone thickness of 3 m. After those first runs, the distance was then fixed and the
thickness varied. These preliminary runs were used to determine the worse cases
structural zone location. Figure 4.6 presents the preliminary soft zone models.

3m
Cv
Cv
3m
Variable

Soft Zone

D T = 10 m D T = 10 m
Variable

Soft Zone

Figure 4.6 – Schematic geometry of preliminary soft zone models

72
Cv

DT = 10 m Thickness
variable

Soft Zone

Figure 4.7 – Schematic geometries of preliminary soft zone models

The next modelling stage was to investigate the variation of the elasticity of a soft
zone according to the host rock mass elasticity. Figure 4.8 shows the layout
geometry used for the elastic models.

73
10 m

Host Rock Cv
Mass

DT
DT
Structural Host Rock
Feature Mass
Structural
Feature

Cv

Host Rock
Mass

Host Rock
Mass
Cv

DT DT
Structural Structural
Feature Feature

Figure 4.8 – Schematic geometries of elasticity models

Orientation of soft zones and hard zones was also investigated. Figure 4.9
schematically illustrates the model geometries.

74
β = 45°

δ = 90° δ = 90°
10 m
120° 120°

Cv
150° 150°
C

180° 180°
DT
DT

45°

Cv

δ = 90° δ = 90°
120° 120°

150° 150°
Cv

180° 180°
DT DT

Figure 4.9 – Schematic geometry of models with variation of orientation of


structural features

4.3.3 Structural anisotropy

Rock mass anisotropy was studied using a FLAC imbedded transversely isotropic
constitutive model. Orientation of anisotropy axis was studied for all the four

75
topographic cases. The anisotropy ratio (E’1/E’2) was studied only for the case of
flat topography. Figure 4.10 shows the details of the anisotropic models.

β = 45o

ω = 90° ω = 90°
60° 60°
Cv
30° 30°
C
0° 0°

DT
DT

β = 52o

Cv
90° β = 52o
60°

90°
30° 60°

30°
0° Cv

DT DT

Figure 4.10 – Schematic geometry of anisotropic models


(angle ω represent the angle of E1)

76
4.3.4 Comparison between models with or without structural features

This section compares hill and valley models with and without soft and hard
structural features. It also introduces the concept of buried valley with the
modelling of overburden filling up the valley. Figures 4.11 and 4.12 are showing
the basic geometry of the comparison models.

β = 52

SOFT STRUCTURAL FEATURE

β = 52

HARD STRUCTURAL FEATURE

β = 52

Figure 4.11 – Schematic geometries of hill comparison models

77
β = 52

SOFT STRUCTURAL FEATURE SOFT STRUCTURAL FEATURE


WITH BURIED VALLEY

β = 52

HARD STRUCTURAL FEATURE HARD STRUCTURAL FEATURE


WITH BURIED VALLEY

β = 52

Figure 4.12 – Schematic geometries of valley comparison models with


overburden

4.3.5 Mohr-Coulomb plastic models

This model uses a Mohr-Coulomb failure criterion with a tension cut-off. This
model involves the yield (failure) of the material resulting in permanent
deformations called plastic flow. The plastic flow formulation assumes that the
total strain increment may be decomposed into elastic and plastic parts, with only
the elastic part controlling stress increments. FLAC proceeds with
implementation, the elastic stress is first computed, and then for each following
step the total strain is incremented. Principal stress is calculated and compared to

78
the criterion. When shear failure happens, a correction must be applied to the
elastic stress to give the new stress state to meet equilibrium in the new
deformation state. Plastic modelling was carried to compare results from elastic
models and to attempt validation. Two sets of modelling runs were done, one with
a soft structural feature surrounded with a hard rock mass, and another with a hard
structural feature surrounded by a soft rock mass. Figure 4.13 shows the
geometry of the plastic models.

β = 45°

δ = 120°
m
10
Cv
C
Structural
Feature
Structural
DT Feature DT

δ = 45°

Cv
δ = 120°
Structural
Feature

δ = 120°

Cv

DT DT
Structural
Feature

Figure 4.13 – Schematic geometries of plastic models

79
4.3.6 Hydro-Mechanical interaction models

Preliminary hydro-mechanical coupling was investigated using FISH, a


programming language embedded in FLAC that allows the user to define new
variables and functions. A computer program was done to link the volumetric
strain with the permeability using the Lombardi (1989a) model (see table 3.7).
The model allows also porosity variation, but at this preliminary stage, porosity
was kept constant in these models. Figure 4.14 illustrates the relationship between
permeability and volumetric strain. The relationship is illustrated by equation 4.2
from Lombardi (1989a), with ε0 the initial volumetric strain and ε the current
volumetric strain. The parameter m was assigned the value of 10 as recommended
in Lombardi (1989a).

m
⎛ ε ⎞
k c = k 0 ⎜⎜ 1 − ⎟⎟ (4.2)
⎝ ε 0 ⎠

4.60E-12

4.10E-12

3.60E-12

3.10E-12
Permeability (m2)

2.60E-12

2.10E-12

1.60E-12

1.10E-12

6.00E-13

1.00E-13
-0.0006 -0.0004 -0.0002 0 0.0002 0.0004 0.0006 0.0008
Strain

Figure 4.14 – Relation between volumetric strain and permeability

80
The model is a simple tunnel with the water table at the surface, constant tunnel
pressure and variation of the rock mass elasticity. Figure 4.15 shows the
schematic geometry of the model type.

Cv

E rm

DT = 6 m

Figure 4.15 – Schematic geometry of the hydro-mechanical


interaction models

The model was intended only for a uniformly fissured, elastic, and saturated rock
mass subject to compressive stress and their variations. In the modelling, all total
stress was kept in compression.

4.3.7 Boundary conditions

In-situ stress

In the parametric study, gravitational stress was studied using flat ground
surfaces, inclined slopes, hills and valley configurations (figure 4.4). In the case
of slopes, hills and valleys, the models were created in stages, in order to simulate
the actual process of erosion as shown in figure 4.16.

81
1. 2.

Figure 4.16 – Assumed in situ stress pattern to model erosion

Figure 4.16 shows how erosion is modelled in two stages:

1. Initial loading; with vertical denoting the gravity stress (σv) and horizontal
the tectonic stress (σh);
2. Erosion unloading; part of the surface is modified to shape an inclined
slope, hill and valley and the tectonic stress remains unchanged;

In the parametric study, initial tectonic stress is modelled for K = 1.

Hydraulic boundaries

Hydraulic conditions have to be defined in models with water interaction. In


FLAC, hydraulic boundary interactions are set for pore pressure and for
saturation. Saturation is defined as the ratio of pore volume occupied by fluid to
total pore volume. Boundaries that are free are impermeable and pore pressure
and saturation can vary according to the flow from nearby zones. Fixed
boundaries are permeable and allow water to leave or enter the grid.

82
Fixed Pore Pressure

Fixed Saturation

Figure 4.17 – Boundary conditions for hydro-mechanical models

So according to figure 4.17, the models have three types of boundaries:

1. The surface boundary, where the pore pressure is fixed and the saturation
is free. So the water is free to flow in or out the surface of the model;
2. The side’s boundaries, where the pore pressure is fixed and the saturation
is fixed. This is the common boundary condition for an applied pore
pressure. Saturation is set to 1, and fluid is exchanged with the outside;
3. The bottom boundary, where the pore pressure is free and the saturation is
free. This is an impermeable boundary and is the default condition. There
is no exchange between the grid and the outside of the model.

4.3.8 Rock mass properties

In the parametric study, two groups of parameters are used: elastic and hydraulic.

83
4.3.8.1 Deformation parameters

Rock mass deformation is the major parameter in this study. The parametric study
uses the model from Nicholson and Bieniawski (1990) presented in table 3.1, as
seen in the following empirical relationship:

Erm = Eir * 2.8X10-5RMR2+0.009e(RMR/22.82) 4.3

Table 4.3 presents the rock mass quality values used in the study. The Poisson
ratio is expected to vary according to rock mass quality, but in this preliminary
modelling it was kept constant.
Table 4.3 - Range of rock mass elastic parameters depending
on rock mass quality
Rock mass GSI1 Intact rock modulus Rock mass modulus Poisson’s
quality category value (%) Eir (GPa) Erm (GPa) ratio
Very good 80 - 100 100 50 0.3
Good 60 – 80 80 26 0.3
Fair 40 – 60 60 12 0.3
Poor 20 – 40 40 5 0.3
Very poor 0 - 20 20 1 0.3
1
GSI is an updated RMR

4.3.8.2 Hydraulic parameters

Rock mass permeability is the second most important parameter used in this
preliminary study. Initial permeability was set constant for all models, despite the
fact that it might be related to the rock mass quality provided by Liu et al. (1999).
However, rock masses with very low quality often present very low permeability.
The link between the GSI and the permeability would therefore only be valid for
non-argillaceous and brittle rocks, and this is why it was set as a constant for this
modelling.

84
4.4 Model parameters

4.4.1 Isotropic models

Input parameters for the modelling with FLAC are summarized in this section.
For isotropic models, the parameters are listed in table 4.4.

Table 4.4 - Isotropic elastic model parameters

Rock mass quality GSI value Density (kN/m3) Bulk modulus (GPa) Shear modulus
category (%) (GPa)
Rock Mass 20 2700 1 0.3
Overburden - 2000 0.001 0.0003
Structural features
Host Rock Mass 75 2700 22 10
Soft Rock Zone 25 2700 4 2

Host Rock Mass 25 2700 4 2


Hard Rock Zone 75 2700 22 10

4.4.2 Transversely isotropic models (anisotropic)

Input parameters for transversely isotropic models are presented in table 4.5. This
model, in which there are two directionally different elastic moduli, gives the
ability to simulate layered elastic rock masses, normal and parallel to the layers,
which are planes of anisotropy.

Table 4.5 - Transversely isotropic model parameters

Rock mass Topography GSI Rock Modulus1 Modulus1 Shear νxy νxz
quality value Mass Ex (GPa) Ey (GPa) modulus
category (%) Density γxy (GPa)
3
(gk/m )
Rock Mass Flat 22 - 50 2700 10.7 5.2 1.2 0.2 0.41
Rock Mass Slope; Hill; 22 - 50 2700 10 2 1.2 0.2 0.41
Valley
1
Ex : Elastic modulus parallel to anisotropy, Ey : Elastic modulus perpendicular to anisotropy

85
4.4.3 Mohr-Coulomb plastic models

Input parameters for the Mohr-Coulomb plastic models are presented in table 4.6.
Soft rock parameters were preliminarily adjusted to reach failure to reasonable
axial displacement (10 - 100 mm).

Table 4.6 - Mohr-Coulomb plastic models parameters

Topo GSI Rock Mass Bulk Shear Friction Cohesion Dilation Tension
value Density Modulus modulus angle (°) (kPa) angle (°) strength
3
(%) (gk/m ) (GPa) γxy (GPa) (kPa)

Soft rock 25 2700 4 2 25 25 8 25


Hard rock 75 2700 22 10 40 1000 25 1000

4.4.4 Hydro-mechanical interaction models

Preliminary hydro-mechanical interaction models were performed in an attempt to


look at the significance of coupling. Only flat topography was modelled with
homogeneous geology. There are four input parameters for hydro-mechanical
interaction models in FLAC (listed in table 4.7). The porosity is a dimensionless
number defined as the ratio of void volume to total volume of an element. For
rock mass, the porosity values are low and might vary from 0.1 to below 0.001
(see table 4.4). The porosity value was fixed for all models at 0.01.

The Itasca Consulting Group (2002) user’s manual gives special recommendation
regarding the use of porosity values lower than 0.2, which is the case for the
modelling in this study. This is because the apparent stiffness of the pore fluid is
proportional to Kw/n. For low values of n, the stiffness may become very large in
comparison to the stiffness of the solid material, causing the FLAC solution to
take a very long time to converge. The manual recommends reducing Kw in this
case. For pure water at room temperature, Kw = 2 X 109 Pa, the value was reduced
to 1 X 106 Pa for this thesis. The permeability required by FLAC is the mobility
coefficient (e.g. units m2). The initial mobility coefficient was set at kp = 1 X 10-12
m2, which is about kd = 1 X 10-6 m/sec in permeability and in the range of
pervious rock masses (see table 3.6).

86
Table 4.7 - Hydraulic parameters in the models

Parameter Values Symbol State


Rock Mass Porosity (%) 0.01 n Fixed
Rock Mass Initial Mobility 10-12 kp Variable
2
coefficient (m )
Bulk Modulus of Water (Pa) 106 Kw Fixed
Density of Water (kg/m3) 1000 ρw Fixed

The mobility coefficient (also called physical permeability) kp is expressed in m2


and is the input required by FLAC, it is related with permeability kd (in m/s) as in:

k pγ w
kd = (4.4)
μ

4.5 Modelling criteria for preliminary tunnel design

4.5.1 Estimation of the effective rock cover

The basic design rule is that nowhere along the unlined pressure tunnel or shaft
should the internal water pressure exceed the minor in-situ principal stress
(Broch, 1984). This design criterion calls for the realization of in-situ stress
measurements somewhere during the project. However, at the beginning of the
projects, it is always necessary to make preliminary estimates of the minimum
stress using some gravity assumptions (as indicated in table 1.2). This is
particularly essential for initial project layout, tunnel alignment and cost
assessment.

The present study introduces the concept of effective cover, where a percentage of
the geometrical cover is subtracted partly due to a disturbance zone around the
tunnel due to stress perturbation.

The objective of this first modelling phase is to study the extent of the disturbance
zones possibly related to the local stress relaxation process, especially for shallow
tunnels. The modelling compares the preliminary estimated cover, done with

87
cover criteria assessments, with the minimum stress value around the estimated
tunnel using numerical models. The study defines a cover alteration ratio being
the minimum stress value around the estimated tunnel using numerical models
divided by the stress given by the preliminary estimated cover. This ratio indicates
the relative agreement of the minimum stress obtained with the numerical models
and the stress obtained by gravity cover rules. This essentially assumes that the
numerical models provided accurate minimum stress and that the criterion is the
estimate. The main objective is to investigate when the criterion is an adequate
estimate to use and under which circumstances.

Numerical models are done to calculate the stress perturbation around the tunnel
and to estimate an equivalent effective rock cover (Ceff). This effective cover
depends on the minimum stress spatial distribution around the tunnel. A
disturbance zone is observed in the models and the reach of that zone above the
crown of the tunnel (measured perpendicular with the ground surface) is called
the cover disturbance (Z). Figure 4.18 illustrates the cover alteration concept and
indicates the location of Ceff , C and disturbance Z (Z = C - Ceff) for a typical flat
ground example.

The cover alteration ratio is illustrated in the equation:

σ min
m
C γ C
CAR = c = eff r = eff (4.5)
σ min Cγ r C

With σmmin the principal minimum stress modelled contour line of the lowest
magnitude that traverses above the tunnel crown without going around it. And
σcmin the minimum stress from cover criterion and is illustrated in the following
equation:

Cγ r cos β
σ min
c
= (4.6)
FS

88
(Pa)
20 m 0.5e6

40 m 1.0e6
C eff

C
60 m 1.5e6

80 m 2.0e6

2.5e6
100 m
Minimum stress 10 m

120 m
isocontours Z Pressure tunnel
C = Criterion estimate; 90 m
C eff = Effective cover from numerical model = 70 m
CAR = C eff / C = 0.78

Figure 4.18 – Minimum stress isocontours and location


of stress alteration from the tunnel

Essentially, this is the equation given by Broch (1984) in table 2.1 and reworked
to isolate the stress and associated with the cover (C) in equation 4.5 using
FS = 1 and γr = 27 kN/m3. In the following modelling study, the density value of
γr = 27 kN/m3 will therefore always be used when converting stress into CAR
with equation 4.4. Furthermore, slope and topographic corrections have always
been carried out when calculating σcmin in equation 4.5.

4.5.2 Horizontal stress

Two types of horizontal stress regime are considered in this section; the initial
input in-situ ratio K and the secondary K’ that takes place after the erosion and/or
the excavation within the model. For all the models K = 1 while K’ spatial
distribution varies. Where K’ < 1, the rock mass is subjected to lateral
decompression. This situation makes gravity criteria not adequate because the

89
minimum available cover is likely to be over-estimated. Investigators like Ming
and Brown (1988) and Alvarez (1997) addressed this issue and recommended
staying outside of K’ < 1 areas. Modelling was performed to outline areas with
K’ < 1 for the elastic cases of hill and valley.

4.5.3 Leakage flow

Gravity criteria do not address rock mass quality effect. One way to preliminarily
check this effect was to model hydro-mechanical interaction, assuming that rock
mass quality is linked to rock mass deformability (see equation 4.2). For hydro-
mechanical models, the primary parameter is the outflow from the tunnel, also
defined as the tunnel leakage in volume of water per unit of time. The model was
set saturated (the water table at ground surface) with a 6 m diameter tunnel and a
vertical cover of 22 m. Therefore, the natural ground water head at the tunnel
center is 25 m. The model introduced 1.25 MPa of internal tunnel pressure,
resulting in a driving pressure of 1.0 MPa. If this value is compared to the
equivalent minimum stress σcmin of 0.6 MPa from equation 4.5, jacking conditions
are anticipated. This situation is intended to enhance the coupling effect. Note that
since models were only pure elastic materials, the possible failure of the rock
mass was not accounted for in the models.

For each model a relative leakage value is obtained, so different leakage values
are obtained with the variation of rock mass deformability parameters. No
absolute leakage values were calculated; the objective was only to compare
relative results in order to find the relationship between the increase in leakage
and rock mass deformation/permeability properties.

90
CHAPTER 5

MODEL PARAMETRIC STUDIES

In this chapter, the results of parametric modelling studies will be described and
discussed. Numerical modelling was used as a tool to assess the limits of
application of the preliminary rock cover criteria and provide guidelines for
preliminary design. Numerical models were developed following the approach
discussed in Chapter 4. The CAR results and models codes and meshes geometry,
summarized in this chapter are provided in the appendices.

5.1 Extent of the effective cover and decompressed areas

As described in chapter 4, the primary objective of the modelling was to verify the
effective cover extension and to locate areas where K’ was below unity. Model
results will provide the limits of application of conventional cover criteria,
especially for cases involving shallow tunnels. The disturbance zone is a function
of the conventional cover (C) minus the effective cover (Ceff). The ratio Ceff/C is
proportional with the cover alteration ratio (CAR) in equation 4.5.

5.1.1 Circular versus D-shape tunnel

In order to simplify the modelling process, circle tunnel sections were modelled.
However, it was verified that the results from circular geometry did not differ
significantly from the use of regular D-shape tunnels, in terms of the Cover
alteration ratio (CAR). Figures 5.1 illustrate the FLAC mesh for the circular
tunnel and figure 5.2 the mesh for the D-shape tunnel. Figure 5.3 shows the
minimum principal stress contours for a circular tunnel and figure 5.4 shows the
minimum principal stress contours for a D-shape tunnel.

91
Figure 5.1 – Typical FLAC mesh for the circular tunnel

Figure 5.2 – Typical FLAC mesh for the D-shape tunnel

92
Figure 5.3 – Typical FLAC minimum principal stress contours
for a circular tunnel

Figure 5.4 – Typical FLAC minimum principal stress contours


for a D-shape tunnel

93
For each circular tunnel diameter and D-shape tunnel equivalent size, a model was
solve, Ceff was located and CAR was calculated. So for a fixed K = 1, several
tunnel geometries were modelled (as presented in figure 4.3). It can be seen on
figure 5.4 that the major difference from figure 5.3 is the stress concentrations
located on both sides of the floor at the intersection with the walls. The modelling
results of circular tunnels versus D-shaped tunnels are presented on figure 5.5.

Figure 5.5 – C/DT ratio versus the difference in the Cover alteration ratio
from the two tunnel geometries

From figure 5.5 it can be seen that the percentage of differences between both
models in terms of CAR is always less than 1.6 %, and if C/DT is higher than 5,
the difference is always less than 0.6 %. For a C/DT smaller than 5, the tunnel size
becomes important compared with the actual rock cover, and the effect of tunnel
geometry appears to take on more importance. Also, for a situation when
geological structure (such as a fault or a dyke) gets close to the tunnel floor, the
D-shape geometry is likely to have important local effects. But despite this, and
since the location of σmmin is predominately located above the crown, the observed

94
differences can be considered negligible and circular tunnels can be modelled and
results are assumed to be representative of both tunnel geometries.

5.1.2 Tunnel diameter versus rock cover and location of K’ <1 areas

As previously discussed, conventional cover criteria do not address the issue of


tunnel diameter versus rock cover. In fact, it is expected that when the tunnel is
close to the surface and has a large diameter, the tunnel itself has a detrimental
influence on the surrounding stress, leading to an increase in the disturbance zone
and possibly to a inaccurate minimum cover estimation using conventional
criteria.

Four types of topography were studied: flat ground, inclined slopes, as well as hill
and valley effects. For all of these topographies, the C/DT ratio was varied and for
each model the disturbance zone was measured as the model result and presented
as the Cover alteration ratio (CAR). So when CAR is below 1, conventional
criteria overestimate the actual cover and when CAR is above 1, conventional
criteria underestimate actual cover. If CAR = 1, this means that the criterion
reflects an accurate answer, with a factor of safety of FS = 1 and γr = 27 kN/m3.

5.1.2.1 Horizontal ground surface

Figure 5.6 illustrates the relationship between the C/DT ratio versus the CAR. The
curve starts with a gentle slope for 15 < C/DT < 50, followed by a transition and
then reveals a sharp decline when C/DT < 5. These results suggest a shallow
tunnel effect, especially when C/DT < 10, which should be considered in tunnel
design.

The highest CAR value is slightly above 0.9; this indicates the minimum tunnel
disturbance effect without any shallow tunnel effect for K = 1. Thus, the factor of
safety would be calculated as FS = 1 / CAR. For a minimum tunnel disturbance
with CAR = 0.9, the factor of safety is 1.1. Therefore, when C/DT < 5, CAR is
around 0.75, which means an FS around 1.3.

95
Figure 5.6 – Cover alteration ratio results for horizontal ground models

5.1.2.2 Inclined slope

Figure 5.7 shows the typical mesh for a 30° slope model and figure 5.8 the
minimum stress contours for 30° slope.

96
Figure 5.7 – Typical FLAC mesh for the 30° slope models

Figure 5.8 – Typical FLAC minimum principal stress contours


for the 30° slope models

97
Figure 5.9 illustrate the relationship between the C/DT ratio versus CAR for all the
slope models. Once again, the curve presents the same relative typical shape.
When the slope is gentle (β = 5°) the results are very close to the flat model
results (see figure 5.6). With the increase in slope (β = 30°and 45°), the CAR
values are all above 1 when C/DT < 5. This seems to indicate the relative
suitability of the cosine correction in the Norwegian criterion.

Figure 5.9 – Cover alteration ratio results for slope models

5.1.2.3 Hill

The problem related with the topographic nose was first addressed by Broch
(1984) and he suggested neglecting these topographic noses which would be, to a
large extent, stress relieved and decompressed. This subject was therefore studied
using numerical modelling to investigate the extent of the actual stress relief zone
and to look for a possible shaving criterion. Figure 5.10 illustrates the topographic
correction concept marked by the dotted line. This figure also explains the

98
parameters used in the modelling: the hill slope (β), the uncorrected cover (C), the
corrected remaining cover (Ccorr), the tunnel driving head (H), the disturbance
zone around the tunnel (Z) and the thickness of the decompressed nose zone (DH).

β
DH
H C

C corr C eff

Figure 5.10 – Typical hill model with correction concept


and important parameters

Figure 5.11, 5.12 and 5.13 are showing the mesh, the minimum stress contours
and the K’ contour respectively for typical hill models.

Figure 5.11 – Typical FLAC mesh for the 52° hill models
99
Figure 5.12 – Typical FLAC minimum principal stress contours
for the 52° hill models

Figure 5.13 – Typical FLAC K’ contours for the 52° hill models (K’ max = 1)

Figure 5.14 illustrates the model results for K’ and the cover disturbance Z. This
information was used to investigate the required topographic correction. The

100
depth of the zone where K’ < 1, is defined as DH. It is striking to observe that DH
extends down to the core of the hill until the slope angle diminishes below 25°.
This observation strongly supports the shaving at the valley level for every
topographic feature with a slope angle above 25°. The figure also presents the
effective radius which is around 20 m for steep hills and increases to 26 m for flat
hills. This effect is probably due to higher stress concentration located underneath
steep and narrow hills.

Table 5.1 presents the results for the six hill slope angles. The last column shows
the resulting effective cover (Ceff) that remain after removing DH and Z to the
cover C (120 m).

β = 64
K<1 K<1 β = 32

β = 52
K<1
β = 25
K<1

D
D

β = 42
K<1 β = 17

D
D

Figure 5.14 – Hill models investigated (DT = 10 m; C = 120 m)

101
Table 5.1 – Results from the hill models (DT = 10 m; C = 120 m)

Slope Angle DH (m) Z (m) Ceff (m)


(deg.)
64 94 21 5
52 87 20 13
42 78 21 21
32 63 22 35
25 35 24 61
17 0 26 94

5.1.2.4 Valley

The crossing of valley bottom was studied by Hartmaier et al. (1998) and they
found that σmin values always gave high factors of safety due to high stress
concentration. They even performed an in-situ measurement program, which
showed minimum stress values about 5 times more than the minimum rock cover
load. Figure 5.15, 5.16 and 5.17 are illustrating the mesh, the minimum stress
contours and the K’ contour respectively for typical valley models.

Figure 5.15 – Typical FLAC mesh for the 52° valley models

102
Figure 5.16 – Typical FLAC minimum principal stress contours s
for the 52° valley models

Figure 5.17 – Typical FLAC K’ contours for the 52° valley models
(K’ max = 1)

Figure 5.18 presents the CAR results from the valley models. It can be observed
that all the models gave a high CAR above 1, which tends to confirm high stress
concentration at the valley bottom. The only area where the CAR is below 1, is
when C/DT < 5 for slope equals or is lower than 25°. The low slope situation (β =
17°) tends very closely toward the flat model results.

103
Figure 5.18 – CAR results from the valley models (DT = 10 m; C = 25 m)

Figure 5.19 illustrates the model results with the decompressed zones where
K’ < 1 and the effective radius extension. So one can observe that the
decompression zones located on the valley walls for steep valley disappear when
the slope decreases below 45°. In addition, table 5.2 gives the numerical results
for the six valley slope angles. In the table, σmmin is the minimum stress used in
equation 4.5 to calculate the CAR. It therefore represents the extent where stress
is disturbed by tunnel proximity. Ceff is the equivalent cover to equal σmmin, with γ
= 27 kN/m3, as defined in equation 4.6, and can be compared to the actual cover

104
(C = 25 m). So, for slope angles greater than 25°, the equivalent cover (Ceff) is
higher than the 25 m, thus putting the Norwegian criterion on the conservative
side. For slopes smaller than 25°, stress concentrations seem to vanish, and the
model behaves very similarly to a flat model.

β = 64
K<1 K<1
β = 32

D D

K<1 β = 52 K<1
β = 25

D D

K<1 β = 42 K<1
β = 17

D D

Figure 5.19 – Valley models investigated (DT = 10 m; C = 25 m)

105
Table 5.2 - Results from the valley models (DT = 10 m; C = 25 m)

Slope Angle (deg.) σmmin (MPa) C’ (m1) Z (m)


64 1.9 70 17
52 1.4 52 15
42 1.0 37 17
32 0.7 26 18
25 0.6 22 14
17 0.4 15 19
1
Equivalent vertical cover required with γ = 27 kN/m3 and σmmin=γC’

5.1.3 Structural features

Several models were done to study the effect of geological structures on the CAR
and the extent of the disturbance zone. The geological structures are of two types,
the soft rock zone, which is basically a zone with a lower modulus of deformation
than the surrounding rock mass, and the hard rock which is defined as a zone with
a higher modulus than the surrounding rock mass. The modelling was done
starting with the study of the distance of a shear zone structure. The next step
involved taking the worst distance (the distance at which the CAR is lower),
where the soft zone thickness was studied. Following this, models were developed
by studying variations in the modulus of the structures. Essentially, the structure
modulus was increased by steps, from a soft zone progressively towards a stiffer
structure.

After taking the worst thickness, soft zone and hard zone orientation were
modelled with the four types of topographies. Rock mass anisotropy was
modelled for the four topographies and anisotropic intensity and orientation were
investigated. Finally, some models were done for hill and valley topographies
incorporating structures (soft zone and hard zone).

5.1.3.1 Distance of a thin soft zone

The first models to include a thin geological structure were done to study the
distance at which the structure creates the lower CAR. This led to a soft zone of

106
3 m thickness being modelled in the vicinity of a 10 m diameter tunnel and a C/DT
ratio of 5. Figure 5.20 shows the two kinds of models produced: the vertical soft
zone and the horizontal soft zone. For the two orientations, the zone was first
located in the middle of the tunnel, and then the zone was moved step by step,
away from the tunnel.

3m
Cv
Cv
3m
Variable

Soft Zone

D T = 10 m D T = 10 m
Variable

Soft Zone

Figure 5.20 – Soft zone modelling scenarios

Figure 5.21 shows the typical minimum principal stress contours for a vertical soft
zone.

Figure 5.21 – Typical FLAC minimum principal stress contours for the
vertical soft zone models

107
Figure 5.22 presents the CAR according to the distance from the tunnel center.
For both shear zone orientations, the lowest CAR values are located from 3 m to 7
m away from the tunnel center. This therefore seems to suggest that the worst
location for a soft zone is close to the tunnel radius (see table 4.5 for modelling
properties). When the distance of the structure increases, the CAR tends towards
the ratio obtained without a structure zone for horizontal ground (see figure 5.6).

Figure 5.22 – Results for the variation of distance of the 3 m


thick soft rock zone (DT = 10 m; C/DT = 5)

5.1.3.2 Thickness of the soft zone

Having the worse distance result, the thickness of the soft zone was varied with
the distance fixed to the tunnel radius. Figure 5.23 presents the geometry of the
thickness variation models. The zone was assumed to be vertical and the thickness
was increased up to 4 times larger than the tunnel diameter. Figure 5.24 shows
CAR results against the diameter to thickness ratio (DT/T). It can be seen that for
108
the different thicknesses, CAR is well below unity and the CAR variations are
small. However, the worst thickness appears to be when T ~ DT. This places
strong emphasis on the detrimental influence of any significant vertical soft
structure. The thickness has a small overall effect on the minimum stress.

Cv

DT = 10 m Thickness
variable

Soft Zone

Figure 5.23 - Variation in the thickness of the soft zone

Figure 5.24 - Results for the variation of thickness of the soft zone

109
5.1.3.3 Variation of elasticity of the structural features

The next series of modelling studied the variation in deformation modulus


between the structure and the host rock mass. Each structure was located at the
side of the tunnel, at a distance of one radius and their thickness was set to T = 10
m, which corresponds to the tunnel diameter. Figure 5.25 illustrates the four
topographies and the structure orientation. For each topography, the E1/E2 ratio
was varied from 0.2 to 10. Where E1 is the rock mass modulus and E2 the
structure modulus. So when E1/E2 < 1, the structure is a hard rock zone, and for
E1/E2 > 1, the structure is a soft rock zone. When E1/E2 =1, there is no structure
in the model, the whole model has the same modulus. Figure 5.26 shows the
minimum principal stress contours for E1/E2 = 0.2, which correspond to a hard
rock tabular feature crossing a soft rock mass.

Figure 5.27 illustrates two different behaviours; the first one is for the flat model,
where CAR is quite stable, with a slight decrease when the soft rock zone
becomes softer. For the three other models, the shape of the curves is more or less
the same; for low E1/E2 values (hard rock zone) CAR decreases rapidly to zero.
The curve reaches a peak near E1/E2 =1 and for higher E1/E2 values (soft rock
zones), CAR decreases, however less sharply than for the hard rock zone. As a
matter of fact, only the slope models show below unity CAR values. For the hill
and valley models, CAR decreases but does not cross the unity value.

Therefore according to these findings, it seems that hard rock structures enclosed
in a softer rock mass provide a more unfavourable situation than soft zone
structures within a harder rock mass. An exception can be found in the flat model,
where the soft rock zones are more unfavourable when they are in flat or slope
geometry.

110
10 m

Host Rock Cv
Mass

DT
DT
Structural Host Rock
Feature Mass
Structural
Feature

Cv

Host Rock
Mass

Host Rock
Mass
Cv

DT DT
Structural Structural
Feature Feature

Figure 5.25 – Layouts of structural feature and different topographies with


variation of elasticity

111
Figure 5.26 – Typical FLAC minimum principal stress contours for
E1/E2 = 0.2 models

Hard structural Soft structural


feature feature

Figure 5.27 - Results for the variation of elasticity of the structural feature
according to rock mass host

112
5.1.3.4 Variation of structural feature orientations (soft zones and hard zones)

This section presents the results of both soft rock zones and hard rock zones, with
variations in the orientation of the structures. Figure 5.28 shows the model
geometries for the four types of topographies. Only the slope topography was not
symmetric and, therefore, this model required an additional structure at 45°.

Soft rock structural features

Figure 5.29 shows the minimum principal stress contours for soft structural
feature models with E1/E2 = 5. Figure 5.30 illustrates the results for the variation
of the orientation of the soft rock zone. It can be seen that for the flat models all of
CAR are again below unity, with vertical structures being the worse. For the other
topographies, it can be observed that for hill and valley, CAR stays above unity,
reflecting possible acceptable minimum stress. Only the slope model shows a
strong decrease in CAR when the angle decreases to 45°. All the lowest CAR
values are present for those situations when the shear zone is perpendicular to the
ground surface.

113
β = 45°

δ = 90° δ = 90°
10 m
120° 120°

Cv
150° 150°
C

180° 180°
DT
DT

45°

Cv

δ = 90° δ = 90°
120° 120°

150° 150°
Cv

180° 180°
DT DT

Figure 5.28 – Variation of structure orientations with different topographies

114
Figure 5.29 – Typical FLAC minimum principal stress contours for soft
structural feature models

Figure 5.30 - Results for the variation of orientation


of the soft rock 10 m thick zones (E1/E2 = 5)

115
Hard rock zone

Figure 5.31 shows the minimum principal stress contours for hard structural
feature models. Figure 5.32 provides the results for the variation in hard structural
feature orientation model with E1/E2 = 0.2. It can be seen that for the flat models
all of the CAR are below unity, just like the shear zone. For the other
topographies, it can be observed that for hills, the CAR stays above unity,
reflecting possible acceptable minimum stress. On the other hand, the slope and
the valley models show a decrease in CAR in the event of an angle decrease. It is
worth noticing the very sharp decrease of the valley model when the angle falls
below 120°. Again here, all the lowest CAR values are when the hard rock zone is
perpendicular to the ground surface.

Figure 5.31 – Typical FLAC minimum principal stress contours for hard
structural feature models

116
Figure 5.32 - Results for the variation of orientation of the hard rock zone

5.1.3.5 Anisotropic models

The anisotropy of the rock mass was investigated using the anisotropy constitutive
model imbedded in FLAC 2D. In this section, the rock mass is modelled as a
transverse isotropic constitutive model. The mechanical properties are reported in
chapter 4, section 4.4.2. All four topography types were modelled with anisotropy
angle variation, in addition to the intensity of the anisotropy being studied on a
flat model. Figure 5.33 illustrates the four topography types with anisotropy angle
variation.

117
β = 45o

ω = 90° ω = 90°
60° 60°
Cv
30° 30°
C
0° 0°

DT
DT

β = 52o

Cv
90° β = 52o
60°

90°
30° 60°

30°
0° Cv

DT DT

Figure 5.33 – Variation of the anisotropy orientations with different


topographies (angle ω represent the angle of E’1)

Figure 5.34 shows the minimum principal stress contours for anisotropic models.
The results of the anisotropic models are presented in figure 5.35. In the case of
flat topography, it appears that the orientation of the angle of anisotropy does not
have a large influence on CAR compared to the other topographic models. This
observation might be related to the fact that flat models are not eroded unlike
other models. No secondary decompression is applied to the model, and CAR
variations are minimized. Furthermore, the effect of anisotropy on the CAR is
exacerbated by the presence of non-flat topographic features (generating irregular

118
model boundaries), compared to the relative regularity of the boundary conditions
in the flat model.

a) ω = 30° b) ω = 30°

c) ω = 0° d) ω = 30°

Figure 5.34 – Typical FLAC minimum principal stress contours for


anisotropic models

Where the other topographies are concerned, variations are more pronounced, and
the CAR value decreases when the principal axis of anisotropy (E’1) is near
horizontal. Figure 5.36 presents the anisotropy intensity variation, which is the
ratio between the principal axis modulus and the secondary axis modulus. This
last model was done with flat topography and 60° anisotropy principal axis angle.
The higher the intensity, the lower the CAR.

119
Figure 5.35 - Results for the variation of the anisotropy angle

Figure 5.36 - Results for the variation of the anisotropy intensity ratio
(E’1/E’2) for flat models

5.1.3.6 Topographic models with structures

Hill

This section compares the same hill topography with or without a vertical
structure. This time the K’ ratio and the disturbance radius (Z) were drawn. Figure
5.37 reveals the model results, with the first having no structure, the second
having a vertical soft zone (E1/E2 = 5) and the third having a vertical hard rock

120
zone (E1/E2 = 0.2). It can be observed that the decompressed zone, where K’ < 1
is similar for all three models. While the disturbance zones are quite different, and
even become very large for the hard rock zone, they remove all possible effective
cover. Table 5.3 provides the numerical values and the effect of the structure is
reflected in the increasing disturbance (Z) and the decreasing effective cover
(Ceff).

β = 52
K<1

SOFT STRUCTURAL FEATURE

β = 52
K<1

HARD STRUCTURAL FEATURE

β = 52
K<1

Figure 5.37 – Comparison between different hill models

121
Table 5.3 - Results from the hill models with
a vertical structure (DT = 10 m; C = 120 m)

Ceff (m) DH (m) Z (m)


Homogeneous 20 85 15
Soft rock zone 25 65 30
Hard rock zone 0 90 90

Valley

Models were also produced to study the effect of overburden filling the valley.
Figure 5.38 illustrates the results of these buried valley models compared with
open valley results. The figure shows that the decompressed zones (K’ < 1)
located on the valley walls vanished when the valley was filled with overburden
with a density of 20 kN/m3. Table 5.4 presents the numerical values and it can be
observed that the overburden slightly increased the minimum stress, therefore the
overburden does increase the available confinement as stated in Eskilsson (1999).
The disturbance zones are quite similar, but have a tendency to increase with the
overburden cover, possibly reflecting a decrease in stress concentration.

122
K<1 β = 52 K<1

SOFT STRUCTURAL FEATURE SOFT STRUCTURAL FEATURE


WITH BURIED VALLEY

K<1 β = 52 K<1

HARD STRUCTURAL FEATURE HARD STRUCTURAL FEATURE


WITH BURIED VALLEY

β = 52
K<1 K<1

Figure 5.38 – K’ and Z results from the buried


valley models (DT = 10 m; C = 25 m)

The presence of vertical structural features at the valley bottom also makes a
significant increase in the disturbance zone extent. Both the shear zone and dyke
more than double the disturbance area for both open and filled valleys.

123
Table 5.4 - Results from the valley models with
a vertical structure (DT = 10 m; C = 25 m)

Open Valley Buried Valley


σmmin (MPa) Z (m) σmmin (MPa) Z (m)
Homogeneous 1.4 10 1.9 12
Soft structural 0.9 20 1.3 25
feature
Hard structural 0.0 25 0.9 46
feature

The minimum stress σmmin from table 5.4 is compared to σcmin = 25 m * 27 kN/m3
= 0.675 MPa that is assumed using the Norwegian criterion.

5.1.3.7 Plastic models with structures

Soft rock zones

Figure 5.39 reveals the minimum stress (σmmin) results for the four topographies
studied with a 10 m thick soft zone. The elastic model is compared to the plastic
model. For all four topographies the minimum stress is greatly decreased with the
plastic model. Table 5.5 presents the comparison between CAR values. With
plastic models the CAR is reduced even for hills and valleys, which presented
high elastic CAR.

In these models, the soft zone fails and greatly deforms especially in the vicinity
of the tunnel. This produces large relaxation of the surrounding hard rock walls.

124
1e5
MPa β = 45
1e5 MPa
5e5 MPa 45 m

Pa
M
0
10 m
37 m

Elastic model
Plastic model
Soft rock zone 10 m

β = 52

β = 52
120 m
Pa
0M
2e

1.1e6
MPa
5
M
Pa

25 m
a
1.4e6 MP

10 m

10 m

Figure 5.39 – Minimum stress results from plastic models compared with
minimum stress from elastic models for soft rock structure

Table 5.5 - CAR results for elastic models compared to plastic models

Topography Elastic CAR Plastic CAR


Flat 0.47 0.08
Slope 0.57 0
Hill 1.17 0.16
Valley 1.61 0

125
Hard rock zone

Figure 5.40 shows the minimum stress (σmmin) results for the four topographies
studied with a 10 m thick hard zone. The elastic model is compared to the plastic
model. In this case stress is increased for the valley model, while for the other
topographies minimum stress is decreased with the plastic model. Table 5.6
presents the comparison between CAR values.

In these models, the soft rock mass fails and greatly deforms, especially in the
vicinity of the hard rock structure and the tunnel. This produces stress
concentration around the tunnel in the valley case, and stress relaxation in the
other topographic cases.

β = 45
1e5 MPa
5e5
45 m MP
a

8e5 MPa Pa
M
0

10 m
37 m

Elastic model
Plastic model
Soft rock zone 10 m

β = 52

β = 52

a 5e5 M
MP Pa 0 MPa
6e5
a
6 MP
1.4e

Figure 5.40 – Minimum stress results from the plastic models compared with
minimum stress from elastic model for hard rock structure

126
Table 5.6 - CAR results for elastic models compared to plastic models

Topography Elastic CAR Plastic CAR


Flat 0.67 0.08
Slope 0.67 0
Hill 1.13 0.49
Valley 0 0.74

5.2 Rock mass quality effect on expected leakage

This section presents a very preliminary attempt to model the coupling effect of
brittle rock mass permeability with the volumetric strain resulting from water
pressure. This modelling is not as extensive as for the preliminary confinement
models. The primary objective was to initially address the effect of rock mass
quality and how it is related with tunnel leakage. The objective was to integrate
the rock mass quality (and therefore its deformability) into a preliminary design
statement.

A total of seven models were produced with different deformation moduli, plus
one uncoupled model for comparison purposes. The output of each model was the
relative leakage flow, which correspond to the difference between the coupled
flow and the uncoupled flow. Figure 5.41 presents the minimum stress for
coupled and uncoupled models and figure 5.42 illustrates the pore pressure for
coupled and uncoupled models. Both figures allow for the evaluation of the
relative effect of permeability coupling. Figure 5.43 presents the change in
-
mobility coefficient distribution compared to a uniform coefficient of 1.0 X 10
12
m2.

127
50 m 0.0

0.25
40
0.5
30

75
0.
20
1.0
10
1.25

0 0 10 20 30 40 50 m
Uncoupled minimum stress (MPa)
Coupled minimum stress (MPa)

Figure 5.41 – Minimum stress for uncoupled and coupled models


50 m
0.2
40
0.4

30
0.6

20 1 .0
0.8

10

00 10 20 30 40 50 m
Uncoupled minimum stress (MPa)
Coupled minimum stress (MPa)

Figure 5.42 – Pore pressure for uncoupled and coupled models

128
50 m

40

4.0
30

0
3.
20
2.0

10
1.0
0 0 10 20 30 40 50 m

Figure 5.43 – Variation of mobility coefficient (X 10-12 m2) compared to a


constant uncoupled coefficient of 1.0 X 10-12 m2

Figure 5.44 presents the relative flow results, which show the difference between
the coupled and uncoupled flow. The data is presented against the deformation
modulus of the rock mass and revealed that when the rock mass is stiff, with
Erm > 15 GPa the coupling has almost no effect. For deformation between 5 to
15 GPa, the coupling presents a small effect, which is lower than 10 % of the
uncoupled flow. However, the coupling effect jumps rapidly, with a significant
increase in the leakage flow, when the deformation modulus is lower than 5 GPa,
which reflects a soft rock mass. Based on these findings, it can be concluded that
coupling is important for rock masses with a deformation modulus lower than
5 GPa; this becomes highly significant for values lower than 2 GPa, which would
reflect a rock mass quality of around RMR < 30%.

129
Figure 5.44 – Coupling model results

5.3 Conclusions and recommendations from the modelling results

Several observations can be made from the numerical modelling results presented
in this chapter. The general conclusions are:

1. Steep hills and valley sides are prone to decompression (K’ < 1).
a. locate the tunnel where K’ > 1

2. The ratio between rock cover (C) and tunnel diameter (DT) significantly
influences the reliability of the Norwegian criterion. The transition
observed around C/DT ~ 10 suggests that an unlined pressure tunnel could
be defined as shallow when C/DT < 10
a. Assure that C/DT > 5

3. Norwegian criterion with the cosine correction is adequate for slope

4. Topographic correction is required when β > 25°


130
a. the correction should be down to the valley bottom elevation

5. Valley bottoms usually contain high stress concentrations making the


Norwegian criterion overly conservative
a. structural stability should be checked against slab buckling
b. valley walls are largely decompressed and contain extensive K’ < 1
areas
c. the presence of thick overburden in valleys have a significant
influence on confinement and on K’ values

6. Structural features
a. Structure location – important effect when d < 2 DT
b. Structure thickness – important effect when T ~ DT
c. Structure orientation – important effect when perpendicular to the
rock surface
d. Anisotropy –important effect when the principal axis of anisotropy
is sub-horizontal

7. Plastic rock mass


a. The presence of a plastic rock mass (like shale, slate and phyllite,
for example) exacerbate the bad influence of a nearby structure
b. Confirms the non adequateness of the Norwegian criterion for such
difficult tunnelling environments

8. Coupling effect significant for low deformation modulus rock mass


a. When Erm < 5 GPa; RMR < 30%

131
CHAPTER 6

PRELIMINARY LAYOUT AND DESIGN GUIDELINES

This chapter compiles the numerical modelling results into integrated design
guidelines for use by practitioners. They are based on the interpretation of all the
information presented in this thesis. Given the vast nature of the subject, projects
are subdivided into three stages:

A. the preliminary layout


B. the investigation
C. the final design

The first contributions to consider are the numerical modelling results, which are
the starting point of the preliminary cover estimation. Also interesting conclusions
are drawn from existing guidelines and case studies and applied to the guidelines.
The preliminary stage requires attention in terms of design guidelines because it
carries a lot important assumptions that will only be verified during investigation
and might have an important economic impact. This is why the proposed
guidelines essentially concentrate on the preliminary stage since it is crucial for
the project.

These guidelines are proposed to fill the constant need for best practice update
and design improvements, but any guideline should always be used with sound
engineering judgment. Guidelines are only helpful if limits are understood, and in
this sense will always be subject to additional improvements and validations. The
guidelines presented for the preliminary layout were partly validated using the
modelling results presented in chapter 5 and, as previously stated, a large
emphasis is given to present recommendations to orient preliminary design. While
the two other stages are still very important and require equal attention,
recommendations and guidelines are given based on the personal experience of
the author, and the work of Merritt (1999), Alvarez (1997), Fernandez (1994),
Ming and Brown (1988), Benson (1987), Brekke and Ripley (1987) amongst
others, as well as information presented in the extensive literature review.

132
6.1 Factor of safety assessment

The project starts with the analysis of regional topographic and geological maps
and possible tunnel alignments. Each alignment should be partitioned; this means
separating zones with similar characteristics and expected hydraulic and
mechanical behaviours. For each zone, unlined tunnel stability must be assessed
using the Norwegian criterion. If the tunnel is identified as unstable, the
alignments must be changed in order to meet with stability criteria, or perhaps an
appropriate permanent liner or semi-liner must be considered. In this manner, the
tunnel alignment is optimized with realistic factor of safety assumptions to
provide solid economic assessments.

Norwegian authors have suggested an FS between 1.1 and 1.3 to be appropriate


for Norwegian rock masses, which are basically hard granitic gneiss. The new
concept of cover alteration ratio was used to study the numerical modelling results
and compare it with the Norwegian assumptions. Numerical models have shown
highly variable cover alteration situations, depending on the topography, geology
and rock mass structural characteristics. It has allowed identifying several adverse
situations where the use of the Norwegian criterion with FS between 1.1 and 1.3
was not adequate. Therefore it is proposed that the factor of safety be adjusted
when adverse geological conditions are prevailing. The CAR concept and the
results from numerical modelling were used to estimate a required factor of safety
during preliminary layout.

6.2 Integration of numerical results into the preliminary layout

The numerical modelling addresses principally the preliminary in-situ stress issue
and their results were integrated into the equation:

FS = Fa * Fb * Fc (6.1)

This relation represents the factor of safety to be used in the Norwegian criterion.
Factor Fa represents the rock mass geological characteristics, Fb addresses the
cover to tunn66

133
el diameter ratio issue and Fc symbolizes the structural geology situation. Table
6.1 shows the proposed values for Fa, Fb and Fc.

Table 6.1 – Proposed FS

Parameter Description Factor of Safety


A. Rock Mass Characteristics Fa
Soft, low modulus (Erm<5GPa) Sedimentary 1.3
Highly anisotropic Metamorphic
Horizontal lying
Medium modulus (5 – 20GPa) Sedimentary 1.2
Anisotropic Metamorphics
Volcanics
Hard, high modulus (>20GPa) Gneiss 1.1
Isotropic, with slight foliation Massive 1.0
B. Cover to tunnel diameter ratio C/DT Fb
<5 Not recommended
5 – 10 1.2
10 – 20 1.15
> 20 1.1
C. Presence of Structures Fc
Within 2 DT of distance Yes 1.1
Consider semi-lined section Perpendicular to surface 1.2
No 1.0

This factor of safety adjustment allows accounting for adverse geological


situations and shallow tunnels. In table 6.1, the possible factors of safety are
between 1.1 and 1.9. So FS carries information on rock mass quality, cover to
tunnel diameter ratio and structural geology.

This factor of safety table is used with the preliminary stability verification table
(table 6.2). It addresses the five leading parameters: rock mass quality,
hydrogeological initial conditions, in-situ stress, permeability, and slope stability.

134
The resulting FS from equation 6.1 includes a correction for two of those
important parameters, the rock mass quality and the minimum stress distribution.

Table 6.2 - Preliminary stability verification table

Parameter Information type


A. Rock mass 1. Regional geology
quality 2. Aerial photographic interpretation
B. Hydrogeology 1. Environmental information (see table 6.3)
2. Assume dry conditions for preliminary estimation
C. In-situ minimum 1. Use of the Norwegian criterion
stress 2. Topographic correction when β > 25°
3. Account for thick overburden
4. Use the Factor of Safety table 6.1 and FS = Fa * Fb *
Fc
D. Permeability 1. Permeability estimation based on figure 3.6 information
2. Calculate leakage using eq. 2.2 and Q < 10 l/s/km
E. Slope stability 1. Verification sliding joint stability (see equation 2.8)

Rock mass quality

According to table 6.2, it can be seen that little information is available


concerning the geology and structural geology of a future project site at the
preliminary stage. Only limited regional information from governmental agencies
exist and can be used. Rock mass quality is assessed based only on broad rock
type, such as granite, schist, and sedimentary.

This parameter is accounted for in table 6.1 with the use of Fa for the rock mass
general characteristics and in Fc with the presence of structural features, such as
fault and dyke.

135
Hydrogeology

Regarding natural pore pressure, table 6.3 presents guidelines to estimate possible
natural pore pressure and assumes a constant drainage scheme for comparison
purposes.

Table 6.3 - Proposed guidelines to evaluate possible effects of natural pore


pressure on the effective driving pressure in tunnels

Type of Natural pore Rock mass Relative effect on


environments pressure or water permeability and tunnel driving
table elevation conductivity pressure
Rainy and temperate High High Large effect
climate High Low Small effect to be neglected
Dry climate Low High Possible local effect
Low Low No effect

Table 6.3 considers initial pore pressure and permeability as the two parameters
responsible for controlling the effectiveness of the natural water pressure that
would counteract the internal pressure from inside the tunnel.

If the natural pore pressure and permeability are high, the media can then be
considered to behave like granular soil and the water table would be well defined
and would directly affect the driving pressure. In essence, the system behaves like
a Darcy type medium. As permeability decreases, head losses increases and the
driving effect is reduced to a certain point. Permeability might decrease in such a
way that capillary effects become important and the flow behaviour would then be
different from a perfect Darcy type medium. This high natural pore pressure
situation is only possible for rainy and temperate climates.

When natural pore pressure is low, which is the case for dry climates, the driving
effect is much lower or perhaps completely absent. However, with high
permeability, local effects could be possible considering heterogeneity in pore
pressure distribution and possible local water aquifers. Finally, when natural pore
pressure and permeability are both low, the driving effect is almost non-existent

136
and should be neglected. Generally for most preliminary projects however, a
conservative approach is taken and rock mass is considered dry.

In-situ minimum stress

As previously stated, for the minimum stress estimation the factor of safety that is
recommended is based on the modelling results. Table 6.2 present the
recommendations to basically use the Norwegian criterion with the topographic
correction to valley level for hill with slopes steeper than 25°, to account for thick
overburden, and to use the FS from equation 6.1.

In an attempt to initially validate the parameters in table 6.1, some case studies
were examined. Table 6.4 presents a list of projects with there actual as built
Norwegian factor of safety compared with the factor of safety calculated with
equation 6.1. This is not a comprehensive list, but it gives a preliminary sense of
the allowable factors to be used.

The estimated FS should be as close as possible to the actual FS. Six out of 13
estimated FS are falling within 0.2 of the actual case and considered equivalent.
Cases with estimated FS higher than the actual FS (cases no. 2, 4, 5, 6, 8),
correspond to projects with unsatisfactory performance without in-situ stress
measurements, or where in-situ measurement were carried following initial filling.
And therefore the FS for those projects should have been higher. On the other
hand, cases where the actual FS is higher than the estimation are both cases
subject to specific situations; high local permeability (cases 9) and unfavourable
geological structures near the surge shaft (cases 11).

So for 11 out of 13 cases, the FS estimation gives a good preliminary design


value. For the two other projects, the FS estimate is too low for other reasons than
in-situ stresses, thus warranting further investigation for permeability and slope
stability.

137
Permeability

At this stage permeability is not known and can only be assessed with textbook
charts such as the one presented in figure 3.6. Therefore, by using equation 2.2
leakage can be estimated assuming reasonable water table elevation, as done in
Rancourt et al. (2006b). This calculation should be done with conservative water
table elevation and using table 3.2.

Slope stability

Slope stability needs to be given final verification in areas where potential rock
wedges may develop, especially when the tunnel approaches the surface. As an
initial estimation, the method from Hoek and Bray (1974) presented in section
2.1.2.2 should be used. For an additional verification, the Illinois criterion (see
table 2.1) should also be satisfied using the reasonable shear strength of the
sliding surface. Structural stability at this stage is only checked where the tunnel
gets close to a valley steep enough to warrant this verification.

138
Table 6.1 - List of projects used to calibrate the factors of safety of table 6.1
Estimation of the Factor of Safety Actual Factor of Safety

Project Reference Geology Fa C/Dt Fb Struc. Fc FS Topo. Actual β H (m) γr FS Performance Remarks
ratio corr. cover (deg.) static Good Not
C (m) good
1. Toulnustouc 1 Rancourt et al. (2006a) Gneiss 1.1 7 1.2 No 1 1.3 No 82.5 10 170 2.7 1.3 X

2. Toulnustouc 2 Rancourt et al. (2006a) Gneiss 1.1 4 1.2 Yes 1.1 1.5 No 50 12 110 2.7 1.2 X Low stress, Drainage
system required
3. Project A Unpublished Granite 1.1 17 1.15 Yes 1.2 1.5 Yes 200 16 330 2.7 1.6 X

4. Project B Unpublished Granite 1.1 14 1.15 Yes 1.2 1.5 Yes 140 16 330 2.7 1.1 X Low stress

5. Case III Swiger (1991) Sedimentary 1.2 19 1.1 No 1 1.3 No 177 30 400 2.7 1.0 X Extensive grouting
Sandstone/shale Drainage system
6. Case IV Swiger (1991) Sedimentary 1.3 10 1.15 No 1 1.5 No 100 12 220 2.7 1.2 X Extensive grouting
Low stress
7. Malana Rancourt et al. (2003) Sedimentary 1.3 >20 1.1 No 1 1.4 Yes 150 30 210 2.6 1.6 X
Quartzite
8. Thibaudeau-Ricard Boyer (2004) Gneiss 1.1 6 1.2 No 1 1.3 No 12 20 60 2.7 0.5 X Extensive repair
and grouting, low C/Dt
9. Lower-Kihansi Dalho et al. (2003) Gneiss 1.2 >20 1.1 Yes 1.1 1.5 No 650 10 850 2.7 2.0 X High permeability
well foliated Local structures
10. Capivari-Cachoeira Bouvard and Pinto (1969) Granite 1.1 >20 1.1 No 1 1.2 No 200 30 400 2.7 1.2 X

11. Mesitas Marulanda and Sedimentary 1.3 >20 1.1 Yes 1.2 1.7 No 200 45 120 2.5 2.9 X Hydraulic jacking
Gutierrez (1999) In the surge shaft
12. North Fork
Stanislaus Schleiss (1988) Mica schist 1.2 >20 1.1 No 1 1.3 No 440 30 730 2.6 1.4 X Good performance

13. Mamquam Hartmaier et al. (1998) Intrusive 1.1 >20 1.1 No 1 1.2 No 110 10 250 2.7 1.2 X Good performance

139
6.3 Investigation

The guidelines for the preliminary layout should only be used in conjunction with
field measurements in the final design. Preliminary guidelines should never
eliminate the need for field measurements.

This section presents several recommendations for investigation work which


should be carried out for every project. Chapter 3 presents a description of
investigations required for each basic parameter shown in table 6.1. This section
is not intended to be a comprehensive text on an investigation for pressure
tunnels. It does not address, for example, other important subjects such as the
location and number of tests required. It should rather be regarded as broad
guidelines to orient the investigation program, which should always be done by
geologists and engineering geologists with large experience in underground civil
structure.

The investigations are developed to answer questions generated from the


preliminary design assessment. Each mechanical zone receives the required
investigation attention, including the expected information and complexity.
Investigation campaigns should always be planned in order to make adjustments
during project development, and to account for unexpected results or problems.

Table 6.5 describes the information needed for each mechanical zone. Zones are
defined as a portion of the tunnel and its surroundings, which can be assumed to
have the same mechanical parameters and should therefore behave in the same
fashion across the entire zone. Given this fact, each zone requires varying
attention, depending on its expected complexity.

140
Table 6.2 - Type of investigation recommended for each zone

Parameter Description Type of work Special Attention


required
A. Topography -Ground surface -Ground elevation -Large boulder
and overburden elevation survey -Buried valley
-Bedrock surface -Geophysical survey
elevation -Boreholes
-Overburden type and
mechanical parameters
B. Rock mass -Rock type with -Geological mapping -Dissolution
mechanical parameters -Geophysical survey -Alteration
-Rock mass -Borehole with -Weathering
classification coring and -Deformability
geocamera -Fault, Dyke
-In-situ testing
-Laboratory testing
C. Hydrogeology -Natural pore pressure -Geophysical survey -Porosity
-Piezometer in -Saturation
borehole
D. In-situ stress -Minimum stress -In-situ stress -Influence of the tunnel
measurements -Existing natural pore
pressure
-Test quality
E. Permeability -Equivalent In-situ permeability -Open joint
permeability measurements - Highly pervious
-Local pervious structure
structure
F. Slope stability -Rock mass stability -Structural mapping -Shear zone
-Joint family -Core orientation -Dyke
characteristics -Joint filling

According to table 6.5, the initial required information from the investigation
concerns the topographic elevation and the overburden nature and thickness in the
vicinity of the tunnel axis. This kind of information is also required around adits
and shafts.

141
The general recommended testing approach was presented in Rancourt and
Mitri (2007) and it can be summarized as follows:

1. All boreholes should be used as often as possible to obtain the most


information from the recovered cores, and tested extensively to extract in-
situ parameters such as natural pore pressure, minimum in-situ stress and
in-situ permeability.
2. A representative number of boreholes should be instrumented with
permanent piezometers.
3. In-situ stress measurements should be planned in two phases: the first
phase is to gather information from surface boreholes and the second
phase would allow validating minimum stress assumptions by testing from
underground adits.
4. In-situ stress measurements would take advantage of tests from different
methods to cross validate numerical results.
5. The same basic procedure should be repeated to assure homogeneity of
results.
6. For stress and permeability measurements, hydraulic jacking and
permeability test procedures could be combined within the same test to
obtain both parameters on the same tested interval.

For in-situ stress measurements, table 6.6 presents proposed guidelines to choose
the best testing methods related to some expected geomechanical conditions. If
the permeability is expected to be high, hydraulic fracturing is not recommended
and hydraulic jacking along with constant pressure tests should be carried out. If
the tested zone is deeply seated with expected high stress, hydraulic jacking and
hydraulic fracturing tests should also be used with possible over-coring methods.

142
Table 6.3 - Proposed testing methods related to expected geomechanical
properties

State of Fracture Permeability Testing Other types of


stress conditions methods measure

Low stress Open and/or closely High Hydraulic Initial


spaced fractures jacking groundwater,
rock mass
Constant
permeability
pressure test
Closed (with Medium Hydraulic Initial
random open) jacking groundwater, rock
and/or largely Constant mass permeability
spaced fractures pressure test
High stress Closely spaced Low Hydraulic Initial
fractures jacking groundwater,
Over-coring
Largely spaced Very low Hydraulic Initial
fractures fracturing groundwater,
Hydraulic Over-coring
jacking

6.4 Final layout

Finally when all investigations are implemented, the information is re-interpreted


in the final design stage. The investigation has allowed locating lined sections and
other semi-lined consolidation. For each zone, the rock mass should be well
known along with the natural pore pressure. Minimum stress is known in all
critical areas such as the end of steel liners, plug locations and any area with
shallow rock cover.

The tunnel may be repartitioned depending on the findings. Stability assessment is


more straightforward due to the availability of information. Nonetheless, unstable
zones at this stage could possibly lead to further investigations in order to narrow
some specific questions, or to tunnel relocation.

143
The guidelines for the length of the steel liner are based on the Norwegian
approach where anywhere along the tunnel the internal pressure (P) should be
smaller than σmin, the minimum measured stress. Consequently, for P > σmin, a
steel liner should be used. At this stage the geotechnical knowledge is high,
minimum stress has been measured and the factor of safety can be in the range of
1.1 to 1.3 with good confidence in field results. The higher the confidence, the
lower the factor of safety.

When P < σmin a steel is not required but a semi-liner may be needed. If the rock
is of poor quality and/or the rock mass is pervious, a semi-liner is likely to be
required. We therefore define two other purposes for semi-liners: the erosion
lining which acts against dissolution, and the seal lining for pervious rock zones.

6.4.1 Erosion lining

Erosion lining is required to assure the basic long term stability of tunnels against
erosion and rock dissolution. Annandale (1996) presented guidelines to classify
erodible natural materials. It can be evaluated in terms of rational correlation
between the rate of energy dissipation of flowing water and the erodibility
classification of the materials. Although the work of Annandale addresses the
erosion downstream of spillways where very high energy dissipation takes place,
it is assumed that rock mass quality is strongly related to the ease of erosion
and/or dissolution. Table 6.7 presents the recommended guidelines for the choice
of erosion liner according to some rock mass parameters.

Table 6.4 – Guidelines for types of erosion lining for estimated rock mass
parameters

Type of lining Soluble Rocks Rock mass parameter


Yes No RMR (%) Q UCS (MPa)
Unlined with rock X > 60 > 20 > 70
bolts
Rock bolts and X 30 – 60 1 – 20 30 – 70
shotcrete
Concrete lined X < 30 <1 < 30

144
Dissolution should be treated independently with geological description and
chemical assays. If soluble rocks are found, special treatments need to be
considered to prevent dissolution during long term operation. Mechanical erosion
of rock is very likely when RMR is below 30 % and UCS below 30 MPa. This
situation usually calls for a concrete liner. With average quality parameters,
shotcrete and rockbolts can be used, and if RMR is above 60 % and UCS above
70 MPa, erosion is not a concern and only rockbolts are required.

6.4.2 Seal lining and allowable gradient

Seal lining is added to prevent leakage from the tunnel and is therefore based on
the permeability and basic quality of the rock mass. Table 6.8 shows the
recommended guidelines given for the seal lining design. The table also provides
guidelines for allowable gradient in the rock mass according to Benson (1989).

Table 6.5 - Guidelines for types of seal lining for estimated rock mass
parameters and permeability

Rock mass RMR Allowable Permeability Type of lining recommended


class (%) hydraulic (cm/s)
gradient
I < 30 <3 k > 10-4 Steel
10-4 < k < 10-6 Steel, membrane
k < 10-6 Reinforced concrete
-4
II 30 – 60 3–8 k > 10 Steel, reinforced concrete with
membrane
10-4 < k < 10-6 Reinforced concrete
k < 10-6 Concrete
III > 60 >8 k > 10-4 Grout and shotcrete
10-4 < k < 10-6 Shotcrete
-6
k < 10 Unlined

Table 6.8 presents three rock mass classes and for each class, three types of
permeability. Allowable gradients are also presented and are based on Benson
(1988). Poor quality rock mass of class I, are soft and deformable and are prone to

145
deformation when pressurized with great effect on the lining. Depending on
permeability, steel liners to reinforced concrete are recommended in this case. For
fair quality rock mass of class II, deformation is less of an issue and permeability
controls the liner requirements. If the rock mass is permeable, steel or membrane
is required, when it is less permeable only a concrete liner is required. For good
rock quality, the high permeable zone is reinforced with grouting and shotcrete.

6.4.3 Slope stability

At this stage of the project, information about the structure geology was obtained
from specific tunnel sections where possible stability issues were identified during
preliminary and investigation stages. As a result, block sliding stability checks
should be made using the Illinois University criterion (see table 2.1). Other
structural instability situations presented in table 3.4 and table 3.5 should be
checked as well, especially with the results from field measurements.

6.4.4 Other considerations

The long term satisfactory behaviour of a project is also dependent on the


following considerations which need to be addressed: design driving pressure, the
water hammer, and operational requirements.

Driving pressure

This is the pressure that acts against the rock mass. It is the pressure developed by
the maximum static water head minus any natural pore pressure already present.
This is the hydraulic pressure used for the design; it is the maximum possible
static water pressure. It should be noted that for the area near the end of the steel
liner or plug, or low cover areas with a high hydraulic gradient, the natural pore
pressure is not present. Consequently, it cannot be used to reduce driving pressure
in these areas. Natural pore pressure is more effective to reduce driving pressure
in areas with high permeability.

146
Water hammer

The water hammer (dynamic pressure) is the sudden increase in pressure


following a rapid drop in flow from the powerhouse. It is a short event that lasts
less than a few minutes and water pressure has no time to propagate deep into the
rock mass. Helwig (1987) presented a theoretical study to estimate the depth to
which significant transient pressures are transferred into the rock mass. It was
found that the effect of the water hammer is limited to a relatively shallow zone
around the tunnel and is the reason why it is not considered in the design pressure.

Operational requirements

Depending on the project, some special requirements are possible, such as the
frequent stopping/starting of a pumped storage project or specific regular
dewatering/filling operation. Pumped storage projects will show a high rate of
recurrence of the dynamic pressure. Those aspects will possibly affect the final
liner design and will necessitate an increase in conservatism during design.
Table 6.9 provides recommendations for filling and the prescribed dewatering rate
according to the rock mass quality.

Table 6.6 - Recommended guidelines for allowable


gradient and filling/dewatering rate

RMR (%) Filling/Dewatering rate (m/h)


< 30 <2
30 – 60 2–5
> 60 5 – 10 (max)

This table is based on values given in Benson (1988) and in Merritt (1999), and
according to this last author, the filling rate should be consistent with the
dewatering rate.

147
CHAPTER 7

CONCLUDING REMARKS

7.1 Numerical modelling study

Numerical modelling was used to study the robustness of the Norwegian criteria
in a variety of topographic settings and geological environments. Simple models
were designed to answer the basic question: “How can we use the Norwegian
criterion taking into account the topography, the rock mass quality and the
presence of structural features?” Models have been crafted to reflect several
natural settings and to study the sensitivity of the results to changes in parameters
and in assumptions.

7.1.1 Spatial variability and parameter estimation

The design of pressure tunnels is a vast subject controlled by several parameters


and physical interactions. A basic problem in the field of geosciences is that the
engineer has to deal with spatial variability and parameter estimation problems.
The engineer is therefore dealing with heterogeneous material and cannot perform
large quantities of testing that would return an acceptable statistical sample. As a
result, a good part of the assessments are based on judgment and past experiences
in order to select the best possible range of design values. The design guidelines
presented in this thesis have tried to address this variability issue by studying
several geological conditions (topographies, geologies and structures) and by
presenting broad recommendations that attempt to include any possible parameter
variation. But there will always be particular and special geological situations.

Tunnel partitioning allows limiting parameter variability and makes the variability
stationary within the same zone. As recommended in this thesis, field efforts
should be made to not only find important geological features, but also less
important features unfavourably oriented with the tunnel. Good partitioning can
then be performed along with related field and laboratory testing.

148
At the preliminary stage, parameters are estimated and factors of safety are
adjusted towards the high side. However, in the final design, parameters are
measured and variability is somehow narrowed and quantified.

As in all the geosciences fields, guidelines represent the framework of a project.


There will always be new projects that will not fit into this frame, however. That
is when the engineer must know that the guidelines have their limitations and are
intended to change and to follow new project technical requirements.

7.1.2 Limitations of numerical modelling

Elastic models

The numerical modelling presented in this study is mostly isotropic, elastic and
continuous in nature. The principal objective of the gravity models was to
estimate the zone of influence of the tunnel within a certain geological setting and
assess the effective resulting rock cover. Also some anisotropic (transversely
isotropic) models were produced, with elastic continuum assumptions. Therefore,
the modelling did not take into account the blocky discontinuous nature of natural
rock masses. The reason for this is that only stress/strain relation was studied and,
consequently, the failure of the rock mass was not addressed.

In this study the continuity assumptions for these models is considered to be


acceptable and still informative for the stress comparisons made in this thesis.
Plastic models were produced to compare results, but further validation is
required.

The models are elastic only and do not consider the important reality of structural
failure (Alvarez, 1997) which might affect hydraulic failure around the tunnel.
This is especially true in the case of a steep valley side with unfavourable jointing,
or with horizontal structural cracks that can lead to the buckling of the valley
floor.

Geological history along with the associated effects on the natural stress pattern
may cause numerical results to depart considerably from the actual stress spatial

149
variation. 3D modelling could also reveal important effect that were not observed
from 2D models.

Plastic models

Some preliminary plastic models were produced to compare with elastic results.
The plastic models have allowed greater deformation and more stress relaxation
than the elastic models. Plastic models might be more realistic to represent soft
and deformable rock masses, such as shale, slate or phyllite.

This modelling has allowed the sensitivity of the CAR to be explored, and results
might represent extreme geological conditions that are not very common but that
are with no doubts extremely unfavourable for any project; further modelling is
required.

Flow models

The numerical flow models have assumed equivalent continuum flow at the scale
of the tunnel. Saturation was also assumed to be 100%, which Norton and Knapp
(1977) revealed is not always the case, especially for good quality rock mass with
non-persistent jointing and unconnected arrays of cracks.

The permeability was linked to the strain used, which is a relation proposed by
Lombardi (1989). The author suggests using the relation only in compression
which was achieved in the modelling. The sole objective of this model was to link
rock mass quality to relative flow, and find a rock mass quality limit where the
flow seems to be greatly influenced by the deformation of the rock mass under
pressurized water.

Equation 4.2 used in this report, is greatly related to the cubic law (eq. 2.3; 3.5).
However, further validation is required.

150
7.2 Design guidelines

Preliminary layout

A great part of this thesis presents the stress estimation problem and attempts to
provide useful preliminary design guidelines. This was done especially for the
Norwegian criterion which is very well known and used extensively. The
Norwegian criterion assumes 2D plane stress conditions, so the topography is
constant in the direction perpendicular to the plane of the studied section. This
would have an important effect especially in mountainous regions, with sharp 3D
topographic variations in all directions.

The proposed factor of safety to be used is subject to specific concerns such as a


complex 3D scenario or large unexpected parameter variations. They are to be
improved with further validation and comparison with existing and future
projects.

Recommendations are made to solely address hydraulic jacking; it does not


address structural and stress tunnel stability. Tunnel shapes other than circular
may have to be analyzed to examine stability. As far as hydraulic jacking is
concerned, the difference between circular and D-shape tunnels is negligible and
thus the recommendations in the current work are equally valid for D-shape
tunnels.

Guidelines should always be used in combination with field measurements in the


final design. Preliminary guidelines never eliminate the need for field
measurements.

Investigation

All geotechnical investigations carried out in the field are subject to large
variations and errors due to several conditions and situations. Confidence in the
measures is obtained with good tests and reproducible results . All results must be
interpreted with great care and be founded on the sound experience and judgment

151
of individuals conducting the study. All field programs are different and should be
adapted from site to site and according to initial project specifications.

Final design

At this stage the geotechnical knowledge is high, minimum stress has been
measured and the factor of safety can be in the range of 1.1 to 1.3 with good
confidence in field results. This factor of safety would be applied to the tunnel
driving pressure at the end of the steel liner, which would be compared to the
minimum stress measured in-situ. Considering the potential implications of
maintenance work in the case of a failure (partial or total) of a pressure tunnel, the
factor of safety should be adjusted with the confidence and the reliability of the
field measures. The more reliable the measures and the reproducible results, the
more confidence the engineer has to fine-tune the factor of safety.

7.3 Summary

This thesis attempted to provide preliminary guidelines for the best initial tunnel
layout and steel liner length. It provided recommendations for each stage of a
project: the preliminary tunnel alignment selection, the investigation and the final
tunnel design. But the main contribution is in the preliminary design stage.

The thesis first presented a broad literature review of past and currently used
design methods and criteria that identified the need for new design guidelines and
analysis rationale. The thesis also discussed the engineering significance of key
parameters controlling leakage losses and long term tunnel stability. These key
parameters were: rock mass quality, hydrogeology, in-situ stress, permeability and
structural stability.

Numerical modelling was used to improve and broaden the scope of the
Norwegian criteria and help strengthen the factor of safety assumptions. To study
the numerical results, a new cover alteration concept was introduced with the
stress/strain analysis. The extent and the distribution of the stress alteration
around the tunnel were compared with the stress obtained from the Norwegian
cover. Some models with hydro-mechanical interactions were attempted to

152
investigate the effect of rock mass quality in its effectiveness to limit leakage. The
permeability of the rock mass was linked to volumetric strain, so for an increase
in water pressure from the tunnel, expansion would take place in the surrounding
rock mass, thus increasing the permeability and leakage. The objective was to find
a range of rock mass quality where coupling becomes significant and account for
it in the preliminary design. The basic results obtained from the numerical
modelling are presented in the following:

1. Steep hills and valley sides are prone to decompression and tunnel should
be located where K’ > 1.
2. The ratio between rock cover (C) and tunnel diameter (DT) influences the
reliability of the Norwegian criterion, and the transition observed around C/DT
~ 10 suggests that an unlined pressure tunnel could be defined as shallow
when C/DT < 10. C/DT should always be kept above 5.
3. Norwegian criterion with the cosine correction is adequate for slope
4. Topographic correction down to the valley bottom is required when β >
25°.
5. Valley bottoms usually contain high stress concentrations making the
Norwegian criterion overly conservative
6. Valley walls are largely decompressed and contain extensive K’ < 1 areas,
but the presence of thick overburden in valleys have a significant influence
on K’ values
7. Structural features
-Structure location – important effect when d < 2 DT
-Structure thickness – important effect when T ~ DT
-Structure orientation – important effect when perpendicular to the
rock surface
-Anisotropy –important effect when the principal axis of anisotropy is
sub-horizontal
8. Coupling effect significant for low deformation modulus rock mass,
especially when Erm < 5 GPa; RMR < 30%

153
Finally new design guidelines were presented for the three main project stages:
the preliminary layout, the investigation and the final design. The main
contribution of the thesis is in the proposal of a new preliminary design approach
based on the use of the Norwegian criterion with the selection of factor of safety
adjusted with adverse geological characteristics. Recommended factor of safety
varies from 1.1 to 1.9 depending on the rock mass geology, the cover versus
tunnel diameter and the structural geology.

For the investigation stage, basic recommendations were presented to obtain the
most useful information from every effort made in the field. Then, final design
guidelines were proposed mostly based on existing practices.

154
7.4 Recommendations for future study

1. Perform 3D and discrete models with structural features which share the
same stress analysis and comparisons with the Norwegian criterion to
study more complex topographical and geological situations.

2. More projects and case studies are needed to increase the database of
factors of safety. This would allow for the validation of the
recommended guidelines with a greater number of projects.

3. Validate the model used for flow analysis and rock mass quality
assessment completed with the results of this thesis. Study detailed
gradient effect around specifics pressure tunnel structures such as;
concrete plugs, liner ends, underground power and adits, etc. This could
be done with the use of discrete codes.

4. Perform more plastic models with the use of a semi-liner in the models
to simulate tunnel support and consolidation. With the objective to
appreciate the tunnel consolidation effect on the CAR.

5. Study natural pore pressure distribution and its effect on the effective
tunnel driving pressure. This could be advantageously done using both
numerical models and field measurements on an actual project.

155
Thesis contribution and originality

The thesis proposes new guidelines to the preliminary design of unlined pressure
tunnels. The specific contributions are:

1. Evaluate by ways of numerical modelling, the effect of surface topography


and geological characteristics of host rock on the minimum principal stress
development around the pressure tunnel.

2. Introduce a cover alteration ratio (CAR) concept to study the stress


disturbance around and above the tunnel with the use of numerical
modelling.

3. Validate the use of the Norwegian criterion under guidelines related with
the rock mass geology, the rock cover versus tunnel diameter and the
structural geology in the tunnel vicinity.

4. Provide guidelines to adjust the factor of safety used in the Norwegian


criterion with the presence of adverse and unfavourable conditions
identified in numerical models.

156
REFERENCES

Abramson, L.W., Lee, T.S., Sharma, S. and G.M. Boyce, 2002, Slope Stability
and Stabilization Methods, 2e ed., Wiley, N.Y., 712 pages.

Alvarez, T.A., 1997. A study of the coupled hydromechanical behavior of jointed


rock masses around pressure tunnels, PhD Thesis, Dep. Of Civil Eng., Univ. of
Illinois at Urbana-Champaign, 648 pages.

Aggson, J.R., Kim, K., 1987, Technical Note - Analysis of Hydraulic Fracturing
Pressure Histories: A Comparison of Five Methods Used to Identify Shut-in
Pressure, Int. J. Rock Mech. & Min. Sci. 24, pp.75-80.

Amadei, B. and Pan, E., 1995, Role of topography and anisotropy when selecting
unlined pressure tunnel alignment, J. of Geotech. Eng., Dec, pp. 879-885.

Annandale, G. W. 2006, Scour Technology, Mechanics and Engineering Practice,


McGraw-Hill, 430 pages. New-York.

Arjang, B., 1989, Pre-mining stresses at some hard rock mines in the Canadian
Shield, Proc. 30th U.S. Symp. On Rock Mech., Tillerson & Wawersiked., pp.
545-551.

Asgian, M., 1989, A numerical model of fluid-flow in deformable naturally


fractured rock masses, Int J Rock Mech Min Sci. 26; pp. 317-328.

Ballivy, G. and Bouja, A., 1992, Problématique des injections de coulis de ciment,
Comptes-rendus, 2e Colloque sur la consolidation et la réfection des
infrastructures par les techniques d’injection, G. Ballivy ed., Univ. Sherbrooke,
May, pp 5-32.

Bandis, S., Lumsden, A.C. and Barton, N., 1983, Fundamentals of rock joint
deformation, Int. J. Rock Mech. & Min. Sci. Geomech. Abstr., 20, pp.249-268.

157
Bart, M., Shao, J.F., Lydzba, D. and Haji-Sotoudeh, M., 2004, Coupled
hydromechanical modeling of rock fractures under normal stress, Can Geotech
J. 41; pp. 686-697.

Barton, N.R., Lien, R. and Lunde, J., 1974, Engineering classification of rock
masses for the design of tunnel support, J. Rock Mech., 6 (4), pp. 189-239.

Barton, N.R., 1986, Deformation phenomena in jointed rock, Géotechnique, 36;


pp. 147-167.

Benson, R.P., 1988, Design of Unlined and Lined Pressure Tunnels, Canadian
Tunneling, 1987/1988, pp. 37-65.

Berg-Christensen, J. and Dannevig, N.T., 1971, Engineering geological


considerations concerning the unlined pressure shaft at the Mauranger Power
Project, Unpublished report, GEOTEAM A/S, Oslo. (in Brekke and Ripley,
1987).

Bieniawski, Z.T., 1973, Engineering classification of jointed rock masses,


Trans S. Afr. Inst. Civ. Engrs. 15, pp.335-344.

Bieniawski, Z.T., 1978, Determining rock mass deformability: Experience from


case histories, Int. J. Rock Mech. & Min. Sci. Geomech. Abstr., 15, pp.237-247.

Biot, M., 1941, General theory of three-dimensional consolidation, Journal of


Applied Physics, 12; pp. 155-164.

Boussinesq, J., 1868, Mémoires sur l’influence des frottements dans les
mouvements réguliers des fluides, J. Math. Pure Appl., Ser. 2, 13, pp.377-424. (in
Tsang and Witherspoon, 1981).

Bouvard, M., Pinto, N., 1969, Aménagement Capivari-Cachoeira, Étude du puits


en charge, La Houille Blanche, 7, pp.747-760.

158
Boyer, L., 2004, Centrale Thibaudeau-Ricard - Travaux exceptionnels de
consolidation post mise en eau, Colloque AEG2004 - Caractérisations et
traitements pour les fondations et les ouvrages d'art, Montréal, pp.1-15.

Brady, B.H.G. and Brown, E.T., 1993, Rock mechanics for underground mining,
2e ed., Chapman & Hall, London, 571 pages.

Brekke, T.L., Bjorlykke, S. and Blindheim, O.T., 1970, Finite element analysis of
the Byrte unlined pressure shaft failure, Large permanent underground openings
(Brekke and Jorstad, ed.). Oslo, Norway: Universitetsforlaget, pp.337-342 (in
Broch, 1988).

Brekke, T.L. and Ripley, B.D., 1987, Design Guidelines for Pressure Tunnels and
Shafts, EPRI, Research Project 1745-17, AP-5273, Final report, 160 pages.

Broch, E., 1982, The Development of unlined pressure shafts and tunnels in
Norway, Rock Mechanics: Cavern and Pressure Shafts, Wittke ed., Balkema,
pp. 545 – 554.

Broch, E., 1984, Unlined High Pressure Tunnels in Areas of Complex


Topography, Water Power & Dam Construction, Nov, pp.21-23.

Broch, E., Dalho, T.S., Hansen, E., 1997, Hydraulic Jacking Tests for Unlined
High Pressure Tunnels, HydroPower'97, pp581-587.

Cappa, F., Guglielmi, Y., Fénart, P., Merrien-Soukatchoff, V. and Thoraval, A.,
2005, Hydromechanical interactions in a fractured carbonate reservoir inferred
from hydraulic and mechanical measurements. Int. J. Rock Mech. &
Min. Sci.; 42, pp. 287-306.

Chen, M., Bai, M., Technical note - Modeling stress-dependent permeability for
anisotropic fractured porous rock, Int. J. Rock Mech. & Min. Sci. 35;
pp. 1113-1119.

159
Cornet, F.H., and Valette, B., 1984, In-situ stress determination from hydraulic
injection test data, J. Geophys. Res., 89, pp.11527-11537.

Cornet, F.H., Li, L., Hulin, J.-P., Ippoloto, I., Kurowski, P., 2003, The
Hydromechanical Behaviour of Fracture: an In-Situ Experimental Case Study,
Int. J. Rock Mech. & Min. Sc., 40, pp.1257-1270.

Corthésy, R., Gill, D.E., Leite, M.H. (1993). "An integrated approach to rock
stress measurement in anisotropic non linear elastic rock", Int.J. Rock Mech. Min.
Sci., Vol. 30, no. 3, pp. 395-411.

Dalho, T., Evans, K.F., Halvorsen, A. et Myrvang, A., 2003, Adverse Effect of
Pore-Pressure Drainage on Stress Measurements Performed in Deep Tunnels: an
Example from the Lower Kihansi Hydroelectric Power Project, Tanzania.,
Int. J. Rock Mech. Min. Sci. & Geomech. Abstr. Vol. 40, pp. 65 – 93.

Dann, H.E., Hertwig, W.P. and Hunter, J.R., 1964, Unlined tunnels of the Snowy
Mountains Hydro-Electric Authority, Australia, ASCE Power Journal, Paper
No. 4071, Oct., pp.47-79.

Deere, D.U., 1983, Unique Geotechnical Problems at Some Hydroelectric


Projects, 7th Panamerican Conf. On Soil Mech. And Foundations Eng.,
Vancouver, pp.865-888.

Doe, T.W., Korbin, G.E., 1987, A Comparison of Hydraulic Fracturing and


Hydraulic Jacking Stress Measurements, 28th US Symp. On Rock Mech., Tuscon,
pp.283-290.

Eskilsson, J.N., 1999, Design of Pressure Tunnels, Proc. Geo-Engineering for


Underground Facilities, Special Publication No. 90, pp.443-458.

Einstein, H. H., 1991, Design and analysis of underground structures in swelling


and squeezing rocks – Chapter 6, in Underground Structures – Design and
Construction, R. S. Sinha Ed., Elsevier, Amsterdam, 516 pages.

160
Ewert, F. K., 1994, Evaluation and interpretation of water pressure tests, Grouting
in the ground, paper 9, Thomas Telford, London, pp. 141-161.

Fernandez, G., 1994, Behavior of Pressure Tunnels and Guidelines for Liner
Design, J. of Geotech. Eng., 120, pp.1768-1789.

Foyo, A., Sanchez, M.A. and C. Tomillo, 2005, A proposal for a Secondary
Permeability Index obtained from water pressure tests in dam foundations,
Engineering Geology, 77, pp. 69-82.

Franklin, J. A. and Chandra, R., 1972, The Slake-Durability test, Int. J. Rock
Mech. & Min. Sci. Vol. 9, pp. 325 – 341.

Freeze, R. A. and Cherry, J. A., 1979, Groundwater, Prentice-Hall, 604 pages.

Gill, D.E., Corthesy, R., Ouellet, J., Dubé, P.B. and Nguyen, D., 1987,
Improvement to standard doorstopper and Leeman cell stress measuring
techniques, Proc. 2nd Int. Symp. On Field Measurements in Rock Mech., Kobe,
Vol. 1, pp. 75-83.

Goodman, R.E., Moye, D.G., van Schalkwyk, A. and Javandel, I., 1965, Ground
water inflows during tunnel driving, Eng. Geol., 2, pp. 39-56.

Goodman, R.E., 1980, Introduction to Rock Mechanics, Wiley & Sons,


New York.

Grimstad, E. and Barton, N., 1993, Updating the Q-system for NMT, Proc. Int.
symp. On sprayed concrete – modern use of wet mix sprayed concrete for
underground support, Fagernes, (eds Kompen, Opsahl and Berg), Oslo,
Norwegian Concrete Assn.

Gronseth, J. M. and Kry, P. R., 1981, Instantaneous shut-in pressure and its
relationship to the minimum stress. Proc. Workshop on Hydraulic Fracture Stress
Measurements, Monterey, CA, pp. 55-60.

161
Guo, F., Morgenstern, N.R., Scott, J.D., 1993, Technical Note - Interpretation of
Hydraulic Fracturing Pressure: A Comparison of Eight Methods Used to Identify
Shut-in Pressure, Int. J. Rock Mech. & Min. Sci. 30, p.627-631.

Haimson, B.C., and Fairhurst, C., 1967, Initiation and extension of hydraulic
fractures in rock, Soc. Petr. Eng. J., 7, pp.310-318.

Haimson, B.C., 1993, The Hydraulic Fracturing Method of Stress Measurement:


Theory and Practice, Stress and Stress Measurements Methods, Chap 14,
Comprehensive Rock Engineering, pp.395-412, Pergamon, Oxford.

Haimson, B.C. and Cornet, F.H., 2003, ISRM Suggested Methods for rock stress
estimation - Part 3: hydraulic fracturing (HF) and/or hydraulic testing of pre-
esisting fractures (HTPF), Int. J. Rock Mech. & Min. Sc., 40, pp.1011-1020.

Hakami, H., 2001, Rock characterisation facility (RCF) shaft sinking - numerical
computations using FLAC, Int. J. Rock Mech. & Min. Sci., 38, pp. 59-65.

Hartmaier, H.H., Doe, T.W. and Dixon, G., 1998, Evaluation of Hydrojacking
Tests for an Unlined Pressure Tunnel, Tunnelling and Underground Space
Technology, 13, pp.393-401.

Helwig, P.C., 1987, A theoretical investigation into the effects of water hammer
pressure surge on rock stability of unlined tunnels, Hydropowers’87, Oslo,
June 22-25, pp. 811-828.

Hoek, E., 1994, Strength of rock masses, ISRM News Journal, 2 (2), pp. 4–16.

Hoek, E., and J. Bray, 1974, Rock Slope Engineering, Institution of Min. Met.,
London.

Hoek, E., and Brown, E.T., 1980, Underground excavations in rock, Institution of
Min. Met., London, 527 pages.

162
Hoek, E., and Brown, E.T., 1997, Practical estimates of rock mass strength, Int. J.
Rock Mech. & Min. Sci. & Geomech. Abstr., 34 (8), p.1165-1186.

Hoek, E. and Diederichs, M. S., 2006, Empirical estimation of rock mass


modulus, Int. J. Rock Mech. & Min. Sci., 43, pp.203-215.

Hoek, E., Kaiser, P.K. and Bawden, W.F., 1995, Support of underground
excavations in hard rock, Balkema, Rotterdam, 215 pages.

Hsieh, P.A. and Neuman, S.P., 1985a, Field Determination of the Three-
Dimensional Hydraulic Conductivity Tensor of Anisotropic Media. 1. Theory,
Water Resources Research, Vol. 21, No. 11, pp. 1655-1665.

Hsieh, P.A., Neuman, S.P., Stiles, G.K. and Simpson, E.S., 1985b, Field
Determination of the Three-Dimensional Hydraulic Conductivity Tensor of
Anisotropic Media. 2. Methodology and Application to Fractured Rocks, Water
Resources Research, Vol. 21, No. 11, pp. 1667-1676.

Hudson, J.A., 1992, Rock Engineering Systems: Theory and Practice, Horwood,
Chicester, 185 pages.

Hudson, J.A., Stephansson, O., Andersson, J., Tsang, C.F. and Jing, L., 2001,
Coupled T-H-M issues relating to radioactive waste repository design and
performance, Int. J. Rock Mech. & Min. Sci., 38, pp. 143-161.

Hvorslev, M. J., 1951, Time lag and soil permeability in groundwater


observations, Bull. No. 36, U.S. Army Engineer Waterways Experiment station,
CE, Vicksburg, Miss. (in Ballivy and Bouja, 1992).

Itasca Consulting Group, 2002, FLAC Fast Lagrangian Analysis of Continua,


User’s Guide, Minneapolis, Ver. 4.0. Online manual: www.itascacg.com.

Jaeger, C. 1961, Rock Mechanics and Hydro-Power Engineering, Part One, Water
Power, Sept., pp. 349-360.

163
James, A. N. and Kirkpatrick, I. M., 1980, Design of foundations of dams
containing soluble rocks and soils, Q. J. Eng. Geol., Vol. 13, pp. 189-198.

James, P., 1991, Stress and Strain during river downcutting, Australian
Geomechanics, pp. 28-31, December.

Jiao, Y. and Hudson, J.A., 1997, The fully-coupled model for Rock Engineering
Systems, Int. J. Rock Mech. & Min. Sci. & Geomech Abstr., 32, pp.491-512.

Junrui, C., Shouyi, L. and Yanqing, W., 2004, Multi-level fracture network model
for coupled seepage and stress fields in rock mass, Commun. Numer. Meth.
Engng., 20, pp. 63-74.

Kranz, R.L, Frankel, A.D., Engelder, T. and Scholz, C.H., 1979, The permeability
of whole and jointed Barre granite, Int. J. Rock Mech. & Min. Sci., 16,
pp. 225-234.

Last, N.C., Harper, T.R., 1989, Response of fractures rock subject to fluid
injection Part I. Development of a numerical model, Tectonophysics, 172; pp. 1-
31.

Leeman, E.R., 1969, The ‘doorstopper’ and triaxial rock stress measuring
instruments developed by the C.S.I.R., J. of the South Afr. Inst. of Min. and
Metal., February, pp.305-339.

Liu, J., Elsworth, D. and Brady, B.H., 1999, Linking stress-dependent effective
porosity and hydraulic conductivity fields to RMR, Int J Rock Mech Min Sci., 36,
pp. 581-596.

Lombardi, G., 1989a, The FES rock mass model – Part 1, Dams Engineering,
Vol. III, Issue 1, pp 49–76.

Lombardi, G., 1989b, The FES rock mass model – Part 2: Some examples, Dams
Engineering, Vol. III, Issue 3, pp 201–221.

164
Lomize, G. M., 1951, Flow in fractured rocks (in Russian), Gosenergoizdat,
Moscow, 127 pages. (in Tsang and Witherspoon, 1981).

Long, J.C.S., Remer, J.S., Wilson, C.R. and Witherspoon, P.A., 1982, Porous
media equivalents for networks of discontinuous fractures, Water Resources
Research, Vol. 18, No. 3, pp. 645-658.

Louis, C., 1969, A study of groundwater flow in jointed rock and its influence on
the stability of rock masses, Rock Mech. Res. Rep. 10, Imp. Coll., London,
90 pages. (in Tsang and Witherspoon, 1981).

Louis, C., 1974, Rock hydraulics, Rock Mechanics - CISM 165, L. Muller ed.,
Springer, pp. 300-387.

Lugeon, M., 1933, Barrages et Géologie, Méthodes de recherches, Terrassement


et imperméabilisation, F. Rouge & Cie, Lausanne, 138 pages.

Martin, C.D. and Chandler, N.A., 1993, Stress heterogeneity and geological
structures, Int. J. Rock Mech. & Min. Sci. 30(7); pp. 993-999.

Marulanda, A, Gutierrez, R., Vallejo, H., 1990, Selection of Equipments for


Hydrofracturing Tests in Permeable Rocks, Proc. Of the Int. Conf. on Mech. of
Jointed and Faulted Rock, Vienna, April, A.A. Balkema, Rotterdam, pp.667-672.

Marulanda, A., Gutierrez, R., 1999, True Behavior of High Pressure Tunnel for
The Guavio Hydroelectric Project, Proc. Geo-Engineering for Underground
Facilities, Special Publication No. 90, pp.1117-1129.

Merritt, A.H., 1999, Geologic and Geotechnical Considerations for Pressure


Tunnel Design, Geotechnical Special Publication, 90, pp.66-81.

Millard, A., Durin, M, Stietel, A., Thoraval, A., Vuillod, E., Baroudi, H., Plas, F.,
Bougnoux, A., Vouille, G., Kobayashi, A., Hara, K., Fujita, T. and Ohnishi, Y.,
1995, Discrete and continuum approaches to simulate the thermo-hydro-

165
mechanical coupling in a large, fractured rock mass, Int. J. Rock Mech. &
Min. Sci. 32(5); pp. 409-434.

Min, K, Rutqvist, J. Tsang, C.F. and Jing, L., 2004, Stress-dependent permeability
of fractured rock masses: a numerical study, Int. J. Rock Mech. & Min. Sc., 41,
pp. 1191-1210.

Ming, L., Brown, E.T., 1988, A Numerical Method for the Analysis of Unlined
Pressure Tunnels in Jointed Rock, Numerical Methods in Geomechanics,
Innsbruck, pp.1473-1480.

Ming, L., Brown, E.T., 1988, Designing Unlined Pressure Tunnels in Jointed
Rock, Water Power & Dam Construction, Nov., pp. 37-41.

Ming, L., 1987, A numerical method for the analysis of unlined pressure tunnels
in jointed rock, PhD thesis, University of London, 319 pages.

Mitri, H.S., Edrissi, R. and Henning, J., 1994, Finite element modelling of cable-
bolted stopes in hard rock underground mines, SME Annual meeting,
Albuquerque, New Mexico, Feb 14-17, no. 94-116.

Nicholson, G. and Bieniawski, Z.T., 1990, Anon-linear deformation modulus


based on rock mass classification, Int. J. Mining & Geol. Eng., vol. 8, pp.181-202.

Noorishad, J., Tsang, C.F., Witherspoon, P.A., 1992, Theoretical and field studies
of coupled hydromechanical behavior of fractured rocks - 1. Development and
verification of a numerical simulator, Int. J. Rock Mech. & Min. Sci. 29;
pp.401-409.

Norton, D. and R. Knapp, 1977, Transport phenomena in hydrothermal systems:


The nature of porosity, American Journal of Science, Vol. 277, October,
pp. 913-936.

Olsson, R., and Brown, N., 1993, An model for hydromechanical coupling during
shearing of rock joints, Int. J. Rock Mech. & Min. Sci., 30 (7), pp. 845-851.

166
Patton, F.D., 1987, Groundwater instrumentation for determining the effect of
minor geologic details on engineering projects, in The art and science of
geotechnical engineering at the dawn of the twenty first century, Prentice-Hall,
Englewood Cliffs, N.J., pp.73-95.

Poiseuille, 1844, Le mouvement des liquides dans les tubes de petits diamètres.

Polubarinova-Kochina, P. Y., 1962, Theory of groundwater movement, translated


from Russian by J.M.R. DeWeist, Princeton Univ. Press, Princeton, 613 pages.
(in Tsang and Witherspoon, 1981).

Pine, R.J. and Batchelor, A.S, 1984, Downward migration of shearing in jointed
rock during hydraulic injections, Int. J. Rock Mech. & Min. Sci., 21(5),
pp. 249-263.

Rahn, W., 1984, Stress Concentration Factors for the Interpretation of


‘Doorstopper’ Stress Measurements in Anisotropic Rocks, Int. J. Rock Mech. &
Min. Sci., 21, pp. 313-326.

Rancourt, A.J., Murphy, D.K., Claisse, M. and Goel, R.P., 2003, Design and
construction of Malana hydroelectric power tunnel, India, RETC Proc., Robinson
R. A. and Marquart, M. ed., SME, Littleton, Co., pp.711-720.

Rancourt, A. J., Murphy, D. K., Whalen, A. and Benson, R., 2006a, Extensive
stress measurements program at the Toulnustouc hydroelectric project – Quebec,
Canada, Proc. of the Int. Symp. on Rock Stress, Trondheim, Norway, June 19-21,
pp 25-33.

Rancourt, A. J., Chartrand, C., Whalen, A. and Bergeron, D., 2006b, Toulnustouc
pressure tunnel leakage estimation, filling, instrumentation and control, 19th
Canadian Tunneling Conf., Vancouver, BC, Canada, Sept., pp. 87-94.

Rancourt, A.J. and Mitri, H. S., 2007, Hydraulic testing for pressure tunnels,
Canadian Tunneling Magazine, pp 32-37.

167
Reinecker, J., Heidbach, O., Tingay, M., Sperner, B. & Müller, B. (2005): The
release 2005 of the World Stress Map (available online at www.world-stress-
map.org).

Ripley, B.D., Brawner, C.O., 1990, Case Histories of the Infuences of Steep
Topography on In-Situ Stresses in Rock, 43rd Can. Geotech. Conf., pp.411-417.

Ripley, B.D., Doe, T.W., Baker, D.G., 1991, Hydraulic Jacking Testing at the
Wahleach Power Tunnel Project, 12th Can. Tunnelling Conf., pp.43-53.

Romm, E.S., 1966, Flow characteristics of fractured rocks (in Russian), Nedra,
Moscow, 283 pages. (in Tsang and Witherspoon, 1981).

Rutqvist, J., Noorishad, J., Chin-Fu, T., Stephansson, O.,1998, Determining of


Fracture Storativity in Hard Rocks using High-Pressure Injection Testing, Water
Resources Res., 34, pp.2551-2560.

Rutqvist, J., Stephansson, O.,1996, A Cyclic Hydraulic Jacking Test to Determine


the in-situ Stress Normal to a Fracture, Int. J. Rock Mech. & Min. Sci. 33,
pp. 695-711.

Schleiss, A.J., 1986, Design of Pervious Pressure Tunnels, Water Power & Dam
Construction, May, pp.21-26.

Schleiss, A., 1988, Design Criteria Applied for the Lower Tunnel of the North
Folk Stanislaus River Hydroelectric Project in California, Rock Mechanics and
Rock Engineering, 21, pp. 161-181.

Serafim, J.L. and Pereira, J.P., 1983, Considerations of the geomechanical


classification of Bienawski, Proc. Int. symp. on engineering geology and
underground constr., Lisbon 1 (II), pp.33-44.

Silvestri, V. and Tabib, C., 1983, Exact determination of gravity stresses in finite
elastic slopes: Part I. Theoretical considerations, Can. Geotech. J., Vol. 20, pp.47-
60.

168
Singhal, B.B.S. and Gupta, R.P., 1999, Applied Hydrogeology of Fractured
Rocks, KluwerAcademic Publishers, London, 400 pages.

Snow, D.T., 1965, A parallel plate model of fractured permeable media, PhD
thesis, Iniv. Of Calif., Berkeley, 331 pages. (in Tsang and Witherspoon, 1981).

Snow, D.T., 1968, Rock fracture spacings, openings, and porosities, J. of the Soil
Mech. Div., ASCE, Jan., pp.73-91.

Swiger, W.F., 1991, Recent Problems in Pressure Tunnels, 3rd Annual


Stanley D. Wilson Mem. Lect., Univ. of Wash., Shannon & Wilson, Inc.
22 pages.

Talobre, J., 1967, Traité de mécanique des Roches, Dunod, Paris, 315 pages.

Terzaghi, K., 1962, Measurement of stresses in rocks, Geotechnique, 12:2, pp.


105-124.

Terzaghi, K., 1963, Stability of steep slopes on hard unweathered rock,


Geotechnique, 12:4, pp. 251-270.

Terzaghi, K., and Peck, R.B. 1967. Soil mechanics in engineering practice, John
Wiley & Sons, Inc., New York. 549 pages.

Van Schalkwyk, A., Jordaan, J. M. and N. Dooge, 1994, Erosion of rock in


unlined spillways, ICOLD, Durban, pp. 555-571.

Wahlstrom, E. E., 1973, Tunneling in rock, Developments in Geotechnical


Engineering 3, Elsevier, Amsterdam, 235 pages.

Walsh, J.B., 1981, Effect of pore pressure and confining pressure on fracture
permeability, Int J Rock Mech Min Sci. 18; pp.429-435.

Wang, H. F., 2000, Theory of linear poroelasticity with applications to


geomechanics and hydrogeology, Princeton Univ. Press. New Jersey. 287 pages.

169
Widmann, R., 1996, International Society for Rock Mechanics Commission on
Rock Grouting, Int. J. Rock Mech. & Min. Sci. 33; pp. 803-847.

Witherspoon, P.A., Amick, C.H., Gale, J.E., Iwai, K., 1979, Observations of a
potential size effect in experimental determination of the hydraulic properties of
fractures, Water Resour Res. 15; pp. 1142-1146.

Zangerl, C., Eberhardt and S. Loew, 2003, Ground settlements above tunnels in
fractured crystalline rock: numerical analysis of coupled hydromechanical
mechanisms, Hydrogeology Journal, 11, pp. 162-173.

Zoback, M.L., 1992, First- and second-order patterns of stress in the lithosphere:
the World Stress Map Project, J. Geophys. Res., 97(B8), pp 11761-11782.

170
APPENDIX A
Model table parameters
PRELIMINARY MODEL GEOMETRY VERIFICATION

Case no. Model Ratio C/DT K DT C σcmin σmmin CAR Deformation modulus Weight
Bulk Shear of rock
(m) (m) (MPa) (MPa) (GPa) (GPa) (kN/m3)

1 Circle 50 1 2 99 2,673 2,53 0,95 1,0 0,3 27


2 Circle 25 1 4 98 2,646 2,42 0,91 1,0 0,3 27
3 Circle 16 1 6 97 2,619 2,32 0,89 1,0 0,3 27
4 Circle 12 1 8 96 2,592 2,24 0,86 1,0 0,3 27
5 Circle 10 1 10 95 2,565 2,17 0,85 1,0 0,3 27
6 Circle 7 1 14 93 2,511 2,04 0,81 1,0 0,3 27
7 Circle 5 1 20 90 2,43 1,88 0,77 1,0 0,3 27
8 Circle 4 1 24 88 2,376 1,79 0,75 1,0 0,3 27
9 Circle 3 1 30 85 2,295 1,67 0,73 1,0 0,3 27
10 Circle 2 1 40 80 2,16 1,48 0,69 1,0 0,3 27
11 Circle 2 1 50 75 2,025 1,29 0,64 1,0 0,3 27
12 D Shape 50 1 2 99 2,673 2,52 0,94 1,0 0,3 27
13 D Shape 25 1 4 98 2,646 2,39 0,90 1,0 0,3 27
14 D Shape 16 1 6 97 2,619 2,29 0,87 1,0 0,3 27
15 D Shape 12 1 8 96 2,592 2,21 0,85 1,0 0,3 27
16 D Shape 10 1 10 95 2,565 2,14 0,83 1,0 0,3 27
17 D Shape 7 1 14 93 2,511 2,01 0,80 1,0 0,3 27
18 D Shape 5 1 20 90 2,43 1,85 0,76 1,0 0,3 27
19 D Shape 4 1 24 88 2,376 1,76 0,74 1,0 0,3 27
20 D Shape 3 1 30 85 2,295 1,64 0,71 1,0 0,3 27
21 D Shape 2 1 40 80 2,16 1,45 0,67 1,0 0,3 27
22 D Shape 2 1 50 75 2,025 1,25 0,62 1,0 0,3 27
ISOTROPIC MODELS

Ratio K DT C c m
Case no. Model Slope C/DT σ min σ min CAR Deformation modulus Weight Weight
Bulk Shear of rock of soils
(m) (m) (MPa) (MPa) (GPa) (GPa) (kN/m3) (kN/m3)
FLAT
1001 isotropic 0 1,5 1 50 75 2,67 2,53 0,95 1 0,3 27
1002 isotropic 0 2 1 40 80 2,65 2,42 0,91 1 0,3 27
1003 isotropic 0 2,8 1 30 84 2,62 2,32 0,89 1 0,3 27
1004 isotropic 0 3,6 1 24 86 2,59 2,24 0,86 1 0,3 27
1005 isotropic 0 4,5 1 20 90 2,57 2,17 0,85 1 0,3 27
1006 isotropic 0 6,6 1 14 92 2,51 2,04 0,81 1 0,3 27
1007 isotropic 0 9,5 1 10 95 2,43 1,88 0,77 1 0,3 27
1008 isotropic 0 12 1 8 96 2,38 1,79 0,75 1 0,3 27
1009 isotropic 0 16 1 6 96 2,30 1,67 0,73 1 0,3 27
1010 isotropic 0 24 1 4 96 2,16 1,48 0,69 1 0,3 27
1011 isotropic 0 49,5 1 2 99 2,03 1,29 0,64 1 0,3 27
SLOPE
2001 isotropic 5 41,8 1 2 84 2,25 2,12 0,94 1 0,3 27
2002 isotropic 5 26,9 1 2 54 1,45 1,34 0,93 1 0,3 27
2003 isotropic 5 13,2 1 4 53 1,42 1,24 0,87 1 0,3 27
2004 isotropic 5 8,6 1 6 52 1,39 1,17 0,84 1 0,3 27
2005 isotropic 5 6,3 1 8 51 1,37 1,11 0,81 1 0,3 27
2006 isotropic 5 5,0 1 10 50 1,34 1,05 0,78 1 0,3 27
2007 isotropic 5 3,4 1 14 48 1,29 0,96 0,75 1 0,3 27
2008 isotropic 5 2,9 1 16 47 1,26 0,92 0,73 1 0,3 27
2009 isotropic 30 27,7 1 2 55 1,30 1,53 1,18 1 0,3 27
2010 isotropic 30 14,7 1 2 29 0,69 0,72 1,05 1 0,3 27
2011 isotropic 30 7,1 1 4 29 0,67 0,66 0,99 1 0,3 27
2012 isotropic 30 4,6 1 6 28 0,65 0,6 0,93 1 0,3 27
2013 isotropic 30 3,4 1 8 27 0,63 0,54 0,86 1 0,3 27
2014 isotropic 30 2,6 1 10 26 0,61 0,49 0,81 1 0,3 27
2015 isotropic 30 1,7 1 14 24 0,57 0,41 0,72 1 0,3 27
2016 isotropic 30 1,5 1 16 23 0,55 0,38 0,70 1 0,3 27
2017 isotropic 45 31,5 1 2 63 1,20 2,09 1,74 1 0,3 27
2018 isotropic 45 20,9 1 2 42 0,80 1,18 1,48 1 0,3 27
2019 isotropic 45 10,3 1 4 41 0,78 1 1,28 1 0,3 27
2020 isotropic 45 6,7 1 6 40 0,77 0,88 1,14 1 0,3 27
2021 isotropic 45 4,9 1 8 40 0,76 0,77 1,02 1 0,3 27
2022 isotropic 45 3,9 1 10 39 0,74 0,68 0,92 1 0,3 27
2023 isotropic 45 2,7 1 14 37 0,72 0,51 0,71 1 0,3 27
2024 isotropic 45 1,8 1 20 35 0,67 0,31 0,46 1 0,3 27
2025 isotropic 45 1,6 1 22 35 0,66 0,26 0,39 1 0,3 27
HILL
3001 isotropic 64 64,5 1 2 129 3,48 1,95 0,56 1 0,3 27
3002 isotropic 64 25,5 1 5 128 3,44 1,81 0,53 1 0,3 27
Ratio K DT C
Case no. Model Slope C/DT σcmin σmmin CAR Deformation modulus Weight Weight
Bulk Shear of rock of soils
(m) (m) (MPa) (MPa) (GPa) (GPa) (kN/m3) (kN/m3)
3003 isotropic 64 12,5 1 10 125 3,38 1,64 0,49 1 0,3 27
3004 isotropic 64 6 1 20 120 3,24 1,34 0,41 1 0,3 27
3005 isotropic 64 4,7 1 25 118 3,17 1,24 0,39 1 0,3 27
3006 isotropic 52 64,5 1 2 129 3,483 2,26 0,65 1 0,3 27
3007 isotropic 52 25,5 1 5 128 3,4425 2,1 0,61 1 0,3 27
3008 isotropic 52 12,5 1 10 125 3,375 1,91 0,57 1 0,3 27
3009 isotropic 52 6 1 20 120 3,24 1,62 0,50 1 0,3 27
3010 isotropic 52 4,7 1 25 118 3,1725 1,52 0,48 1 0,3 27
3011 isotropic 42 64,5 1 2 129 3,483 2,49 0,71 1 0,3 27
3012 isotropic 42 25,5 1 5 128 3,4425 2,31 0,67 1 0,3 27
3013 isotropic 42 12,5 1 10 125 3,375 2,11 0,63 1 0,3 27
3014 isotropic 42 6 1 20 120 3,24 1,82 0,56 1 0,3 27
3015 isotropic 42 4,7 1 25 118 3,186 1,72 0,54 1 0,3 27
3016 isotropic 32 64,5 1 2 129 3,483 2,73 0,78 1 0,3 27
3017 isotropic 32 25,5 1 5 128 3,4425 2,55 0,74 1 0,3 27
3018 isotropic 32 12,5 1 10 125 3,375 2,32 0,69 1 0,3 27
3019 isotropic 32 6 1 20 120 3,24 2,01 0,62 1 0,3 27
3020 isotropic 32 4,7 1 25 118 3,186 1,72 0,54 1 0,3 27
3021 isotropic 25 64,5 1 2 129 3,483 2,91 0,84 1 0,3 27
3022 isotropic 25 25,5 1 5 128 3,4425 2,71 0,79 1 0,3 27
3023 isotropic 25 12,5 1 10 125 3,375 2,47 0,73 1 0,3 27
3024 isotropic 25 6 1 20 120 3,24 2,15 0,66 1 0,3 27
3025 isotropic 25 4,7 1 25 118 3,1725 2,05 0,65 1 0,3 27
3026 isotropic 17 64,5 1 2 129 3,483 3,05 0,88 1 0,3 27
3027 isotropic 17 25,5 1 5 128 3,4425 2,85 0,83 1 0,3 27
3028 isotropic 17 12,5 1 10 125 3,375 2,6 0,77 1 0,3 27
3029 isotropic 17 6 1 20 120 3,24 2,26 0,70 1 0,3 27
3030 isotropic 17 4,7 1 25 118 3,186 2,16 0,68 1 0,3 27

VALLEY
4001 isotropic 64 14,5 1 2 29 0,783 2,38 3,04 1 0,3 27
4002 isotropic 64 5,5 1 5 28 0,7425 2,23 3,00 1 0,3 27
4003 isotropic 64 2,5 1 10 25 0,675 1,97 2,92 1 0,3 27
4004 isotropic 64 1,5 1 15 23 0,6075 1,73 2,85 1 0,3 27
4005 isotropic 52 14,5 1 2 29 0,783 1,91 2,44 1 0,3 27
4006 isotropic 52 5,5 1 5 28 0,7425 1,63 2,20 1 0,3 27
4007 isotropic 52 2,5 1 10 25 0,675 1,43 2,12 1 0,3 27
4008 isotropic 52 1,5 1 15 23 0,6075 1,25 2,06 1 0,3 27
4009 isotropic 42 14,5 1 2 29 0,783 1,43 1,83 1 0,3 27
4010 isotropic 42 5,5 1 5 28 0,7425 1,28 1,72 1 0,3 27
4011 isotropic 42 2,5 1 10 25 0,675 1,09 1,61 1 0,3 27
4012 isotropic 42 1,5 1 15 23 0,6075 0,94 1,55 1 0,3 27
Ratio K DT C
Case no. Model Slope C/DT σcmin σmmin CAR Deformation modulus Weight Weight
Bulk Shear of rock of soils
(m) (m) (MPa) (MPa) (GPa) (GPa) (kN/m3) (kN/m3)
4013 isotropic 32 14,5 1 2 29 0,783 1,12 1,43 1 0,3 27
4014 isotropic 32 5,5 1 5 28 0,7425 0,97 1,31 1 0,3 27
4015 isotropic 32 2,5 1 10 25 0,675 0,79 1,17 1 0,3 27
4016 isotropic 32 1,5 1 15 23 0,6075 0,65 1,07 1 0,3 27
4017 isotropic 25 14,5 1 2 29 0,783 0,93 1,19 1 0,3 27
4018 isotropic 25 5,5 1 5 28 0,7425 0,79 1,06 1 0,3 27
4019 isotropic 25 2,5 1 10 25 0,675 0,62 0,92 1 0,3 27
4020 isotropic 25 1,5 1 15 23 0,6075 0,48 0,79 1 0,3 27
4021 isotropic 17 14,5 1 2 29 0,783 0,81 1,03 1 0,3 27
4022 isotropic 17 5,5 1 5 28 0,7425 0,66 0,89 1 0,3 27
4023 isotropic 17 2,5 1 10 25 0,675 0,5 0,74 1 0,3 27
4024 isotropic 17 1,5 1 15 23 0,6075 0,38 0,63 1 0,3 27
4025 isotropic with buried valley 52 2,5 1 10 25 0,6075 1,9 3,13 1 0,3 27 20
ANISOTROPIC MODELS

Case no. Model Slope Ratio K DT C σcmin σmmin CAR Anisotropy E1 E2 G12 Weight nuy nuz
C/DT angle intensity // to aniso. -/ to aniso. Rock
(GPa) (GPa) (GPa) (kN/m3)
FLAT
1012 anisotropic 0 4,5 1 10 45 1,215 0,68 0,56 0 2 10,7 5,2 1,2 27 0,2 0,41
1013 anisotropic 0 4,5 1 10 45 1,215 0,73 0,60 30 2 10,7 5,2 1,2 27 0,2 0,41
1014 anisotropic 0 4,5 1 10 45 1,215 0,69 0,57 60 2 10,7 5,2 1,2 27 0,2 0,41
1015 anisotropic 0 4,5 1 10 45 1,215 0,64 0,53 90 2 10,7 5,2 1,2 27 0,2 0,41
1016 anisotropic 0 4,5 1 10 45 1,215 0,83 0,68 60 1 10,7 5,2 1,2 27 0,2 0,41
1017 anisotropic 0 4,5 1 10 45 1,215 0,73 0,60 60 1,3 10,7 5,2 1,2 27 0,2 0,41
1018 anisotropic 0 4,5 1 10 45 1,215 0,71 0,58 60 1,5 10,7 5,2 1,2 27 0,2 0,41
1019 anisotropic 0 4,5 1 10 45 1,215 0,69 0,57 60 1,8 10,7 5,2 1,2 27 0,2 0,41
1020 anisotropic 0 4,5 1 10 45 1,215 0,67 0,55 60 2 10,7 5,2 1,2 27 0,2 0,41
1021 anisotropic 0 4,5 1 10 45 1,215 0,53 0,44 60 4 10,7 5,2 1,2 27 0,2 0,41
SLOPE
2035 anisotropic 45 3,4 1 10 34 0,706 0,10 0,14 0 5 10,0 2,0 1,2 27 0,2 0,41
2036 anisotropic 45 3,4 1 10 34 0,706 0,30 0,42 30 5 10,0 2,0 1,2 27 0,2 0,41
2037 anisotropic 45 3,4 1 10 34 0,706 0,50 0,71 60 5 10,0 2,0 1,2 27 0,2 0,41
2038 anisotropic 45 3,4 1 10 34 0,706 0,70 0,99 90 5 10,0 2,0 1,2 27 0,2 0,41
2039 anisotropic 45 3,4 1 10 34 0,706 0,90 1,27 135 5 10,0 2,0 1,2 27 0,2 0,41
HILL
3031 anisotropic 1 6,7 1 6 40,2 1,674 0,3 0,179 0 5 10,0 2,0 1,2 27 0,2 0,41
3032 anisotropic 1 6,7 1 6 40,2 1,674 0,7 0,418 30 5 10,0 2,0 1,2 27 0,2 0,41
3033 anisotropic 1 6,7 1 6 40,2 1,674 0,9 0,538 60 5 10,0 2,0 1,2 27 0,2 0,41
3034 anisotropic 1 6,7 1 6 40,2 1,674 1 0,597 90 5 10,0 2,0 1,2 27 0,2 0,41
VALLEY
4026 anisotropic 1 2 1 6 12 0,324 0,4 1,235 0 5 10,0 2,0 1,2 27 0,2 0,41
4027 anisotropic 1 2 1 6 12 0,324 0,3 0,926 30 5 10,0 2,0 1,2 27 0,2 0,41
4028 anisotropic 1 2 1 6 12 0,324 0,5 1,543 60 5 10,0 2,0 1,2 27 0,2 0,41
4029 anisotropic 1 2 1 6 12 0,324 0,9 2,778 90 5 10,0 2,0 1,2 27 0,2 0,41
MODELS WITH STRUCTURAL FEATURES

Ratio DT C
Case noModel Angle σcmin σmmin CAR Structural feature Rock mass Weight Weight
of slope C/DT Thickness Angle Dist. from Bulk Shear Bulk Shear Rock of soils
tunnel center
(deg.) (m) (m) (MPa) (MPa) (m) (deg) (m) (GPa) (GPa) (GPa) (GPa) (kN/m3) (kN/m3)
FLAT
1022 isotropic with a vertical soft zone 0 4,5 10 45 1,215 0,79 0,6502 3 90 0 4 2 22 10 27
1023 isotropic with a vertical soft zone 0 4,5 10 45 1,215 0,72 0,5926 3 90 3 4 2 22 10 27
1024 isotropic with a vertical soft zone 0 4,5 10 45 1,215 0,63 0,5185 3 90 5 4 2 22 10 27
1025 isotropic with a vertical soft zone 0 4,5 10 45 1,215 0,57 0,4691 3 90 7 4 2 22 10 27
1026 isotropic with a vertical soft zone 0 4,5 10 45 1,215 0,86 0,7078 3 90 10 4 2 22 10 27
1027 isotropic with a vertical soft zone 0 4,5 10 45 1,215 0,85 0,6996 3 90 15 4 2 22 10 27
1028 isotropic with a vertical soft zone 0 4,5 10 45 1,215 0,84 0,6914 3 90 20 4 2 22 10 27
1029 isotropic with a vertical soft zone 0 4,5 10 45 1,215 0,83 0,6831 3 90 25 4 2 22 10 27
1030 isotropic with a vertical soft zone 0 4,5 10 45 1,215 0,83 0,6831 3 90 30 4 2 22 10 27
1031 isotropic with a horizontal soft zone 0 4,5 10 45 1,215 0,8 0,6584 3 0 0 4 2 22 10 27
1032 isotropic with a horizontal soft zone 0 4,5 10 45 1,215 0,73 0,6008 3 0 3 4 2 22 10 27
1033 isotropic with a horizontal soft zone 0 4,5 10 45 1,215 0,75 0,6173 3 0 5 4 2 22 10 27
1034 isotropic with a horizontal soft zone 0 4,5 10 45 1,215 0,77 0,6337 3 0 7 4 2 22 10 27
1035 isotropic with a horizontal soft zone 0 4,5 10 45 1,215 0,82 0,6749 3 0 10 4 2 22 10 27
1036 isotropic with a horizontal soft zone 0 4,5 10 45 1,215 0,87 0,716 3 0 15 4 2 22 10 27
1037 isotropic with a horizontal soft zone 0 4,5 10 45 1,215 0,85 0,6996 3 0 20 4 2 22 10 27
1038 isotropic with a horizontal soft zone 0 4,5 10 45 1,215 0,84 0,6914 3 0 25 4 2 22 10 27
1039 isotropic with a horizontal soft zone 0 4,5 10 45 1,215 0,83 0,6831 3 0 30 4 2 22 10 27
1040 isotropic with a soft zone oriented 0 deg. 0 4,5 10 45 1,215 0,71 0,5844 3 0 5 4 2 22 10 27
1041 isotropic with a soft zone oriented 30 deg. 0 4,5 10 45 1,215 0,71 0,5844 3 30 5 4 2 22 10 27
1042 isotropic with a soft zone oriented 60 deg. 0 4,5 10 45 1,215 0,62 0,5103 3 60 5 4 2 22 10 27
1043 isotropic with a soft zone oriented 90 deg. 0 4,5 10 45 1,215 0,62 0,5103 3 90 5 4 2 22 10 27
1044 isotropic with a soft zone oriented 0 deg. 0 4,5 10 45 1,215 0,71 0,5844 10 0 5 4 2 22 10 27
1045 isotropic with a soft zone oriented 30 deg. 0 4,5 10 45 1,215 0,68 0,5597 10 30 5 4 2 22 10 27
1046 isotropic with a soft zone oriented 60 deg. 0 4,5 10 45 1,215 0,57 0,4691 10 60 5 4 2 22 10 27
1047 isotropic with a soft zone oriented 90 deg. 0 4,5 10 45 1,215 0,57 0,4691 10 90 5 4 2 22 10 27
1048 isotropic with a soft zone oriented 0 deg. 0 4,5 10 45 1,215 0,79 0,6502 inf 0 5 4 2 22 10 27
1049 isotropic with a soft zone oriented 30 deg. 0 4,5 10 45 1,215 0,73 0,6008 inf 30 5 4 2 22 10 27
1050 isotropic with a soft zone oriented 60 deg. 0 4,5 10 45 1,215 0,6 0,4938 inf 60 5 4 2 22 10 27
1051 isotropic with a soft zone oriented 90 deg. 0 4,5 10 45 1,215 0,57 0,4691 inf 90 5 4 2 22 10 27
1052 isotropic with a hard zone oriented 0 deg. 0 4,5 10 45 1,215 0,87 0,716 3 0 5 22 10 4 2 27
1053 isotropic with a hard zone oriented 30 deg. 0 4,5 10 45 1,215 0,51 0,4198 3 30 5 22 10 4 2 27
1054 isotropic with a hard zone oriented 60 deg. 0 4,5 10 45 1,215 0,5 0,4115 3 60 5 22 10 4 2 27
1055 isotropic with a hard zone oriented 90 deg. 0 4,5 10 45 1,215 0,86 0,7078 3 90 5 22 10 4 2 27
1056 isotropic with a hard zone oriented 0 deg. 0 4,5 10 45 1,215 0,91 0,749 10 0 5 22 10 4 2 27
1057 isotropic with a hard zone oriented 30 deg. 0 4,5 10 45 1,215 0,85 0,6996 10 30 5 22 10 4 2 27
1058 isotropic with a hard zone oriented 60 deg. 0 4,5 10 45 1,215 0,83 0,6831 10 60 5 22 10 4 2 27
1059 isotropic with a hard zone oriented 90 deg. 0 4,5 10 45 1,215 0,82 0,6749 10 90 5 22 10 4 2 27
1060 isotropic with a hard zone oriented 0 deg. 0 4,5 10 45 1,215 0,88 0,7243 inf 0 5 22 10 4 2 27
1061 isotropic with a hard zone oriented 30 deg. 0 4,5 10 45 1,215 0,79 0,6502 inf 30 5 22 10 4 2 27
FLAT
Ratio DT C
Case noModel Angle σcmin σmmin CAR Structural feature Rock mass Weight Weight
of slope C/DT Thickness Angle Dist. from Bulk Shear Bulk Shear Rock of soils
tunnel center
(deg.) (m) (m) (MPa) (MPa) (m) (deg) (m) (GPa) (GPa) (GPa) (GPa) (kN/m3) (kN/m3)
1062 isotropic with a hard zone oriented 60 deg. 0 4,5 10 45 1,215 0,86 0,7078 inf 60 5 22 10 4 2 27
1063 isotropic with a hard zone oriented 90 deg. 0 4,5 10 45 1,215 0,79 0,6502 inf 90 5 22 10 4 2 27
1064 isotropic with a soft zone with varying thickness 0 4,5 10 45 1,215 0,63 0,5185 2 90 5 4 2 22 10 27
1065 isotropic with a soft zone with varying thickness 0 4,5 10 45 1,215 0,62 0,5103 3 90 5 4 2 22 10 27
1066 isotropic with a soft zone with varying thickness 0 4,5 10 45 1,215 0,59 0,4856 5 90 5 4 2 22 10 27
1067 isotropic with a soft zone with varying thickness 0 4,5 10 45 1,215 0,57 0,4691 10 90 5 4 2 22 10 27
1068 isotropic with a soft zone with varying thickness 0 4,5 10 45 1,215 0,58 0,4774 15 90 5 4 2 22 10 27
1069 isotropic with a soft zone with varying thickness 0 4,5 10 45 1,215 0,59 0,4856 20 90 5 4 2 22 10 27
1070 isotropic with a soft zone with varying thickness 0 4,5 10 45 1,215 0,59 0,4856 25 90 5 4 2 22 10 27
1071 isotropic with a soft zone with varying thickness 0 4,5 10 45 1,215 0,59 0,4856 inf 90 5 4 2 22 10 27
1072 isotropic with a variation in zone elastic parameters 0 4,5 10 45 1,215 0,83 0,6831 10 90 5 4 2 22 10 27
1073 isotropic with a variation in zone elastic parameters 0 4,5 10 45 1,215 0,83 0,6831 10 90 5 4 2 10 5 27
1074 isotropic with a variation in zone elastic parameters 0 4,5 10 45 1,215 0,84 0,6914 10 90 5 4 2 7 3 27
1075 isotropic with a variation in zone elastic parameters 0 4,5 10 45 1,215 0,84 0,6914 10 90 5 4 2 5 2,5 27
1076 isotropic with a variation in zone elastic parameters 0 4,5 10 45 1,215 0,84 0,6914 10 90 5 9 4 9 4 27
1077 isotropic with a variation in zone elastic parameters 0 4,5 10 45 1,215 0,72 0,5926 10 90 5 10 5 22 10 27
1078 isotropic with a variation in zone elastic parameters 0 4,5 10 45 1,215 0,64 0,5267 10 90 5 7 3 22 10 27
1079 isotropic with a variation in zone elastic parameters 0 4,5 10 45 1,215 0,58 0,4774 10 90 5 4 2 22 10 27
1080 isotropic with a variation in zone elastic parameters 0 4,5 10 45 1,215 0,58 0,4774 10 90 5 2 1 22 10 27
SLOPE
2043 isotropic with a soft zone 45 4 10 40 0,7064 0,68 0,9626 10 homo 5 4 2 22 10 27
2044 isotropic with a soft zone 45 4 10 40 0,7064 0,89 1,2599 10 0 5 4 2 22 10 27
2045 isotropic with a soft zone 45 4 10 40 0,7064 0,66 0,9343 10 30 5 4 2 22 10 27
2046 isotropic with a soft zone 45 4 10 40 0,7064 0,85 1,2033 10 60 5 4 2 22 10 27
2047 isotropic with a soft zone 45 4 10 40 0,7064 0,63 0,8918 10 90 5 4 2 22 10 27
2048 isotropic with a soft zone 45 4 10 40 0,7064 0,4 0,5663 10 135 5 4 2 22 10 27
2049 isotropic with a hard zone 45 4 10 40 0,7064 0,68 0,9626 10 homo 5 22 10 4 2 27
2050 isotropic with a hard zone 45 4 10 40 0,7064 0,82 1,1608 10 0 5 22 10 4 2 27
2051 isotropic with a hard zone 45 4 10 40 0,7064 0,91 1,2882 10 30 5 22 10 4 2 27
2052 isotropic with a hard zone 45 4 10 40 0,7064 0,77 1,09 10 60 5 22 10 4 2 27
2053 isotropic with a hard zone 45 4 10 40 0,7064 0,41 0,5804 10 90 5 22 10 4 2 27
2054 isotropic with a hard zone 45 4 10 40 0,7064 0,47 0,6653 10 135 5 22 10 4 2 27
2055 isotropic with a variation in zone elastic parameters 45 4 10 40 0,7064 0,001 0,0014 10 90 5 4 2 22 10 27
2056 isotropic with a variation in zone elastic parameters 45 4 10 40 0,7064 0,03 0,0425 10 90 5 4 2 10 5 27
2057 isotropic with a variation in zone elastic parameters 45 4 10 40 0,7064 0,51 0,722 10 90 5 4 2 7 3 27
2058 isotropic with a variation in zone elastic parameters 45 4 10 40 0,7064 0,8 1,1325 10 90 5 4 2 5 2,5 27
2059 isotropic with a variation in zone elastic parameters 45 4 10 40 0,7064 0,86 1,2174 10 90 5 9 4 9 4 27
2060 isotropic with a variation in zone elastic parameters 45 4 10 40 0,7064 0,68 0,9626 10 90 5 10 5 22 10 27
2061 isotropic with a variation in zone elastic parameters 45 4 10 40 0,7064 0,55 0,7786 10 90 5 7 3 22 10 27
2062 isotropic with a variation in zone elastic parameters 45 4 10 40 0,7064 0,39 0,5521 10 90 5 4 2 22 10 27
2063 isotropic with a variation in zone elastic parameters 45 4 10 40 0,7064 0,23 0,3256 10 90 5 2 1 22 10 27

HILL
3037 isotropic with a soft zone 52 12,5 10 125 3,375 1,65 0,4889 10 0 5 4 2 22 10 27
Ratio DT C
Case noModel Angle σcmin σmmin CAR Structural feature Rock mass Weight Weight
of slope C/DT Thickness Angle Dist. from Bulk Shear Bulk Shear Rock of soils
tunnel center
(deg.) (m) (m) (MPa) (MPa) (m) (deg) (m) (GPa) (GPa) (GPa) (GPa) (kN/m3) (kN/m3)
3038 isotropic with a soft zone 52 12,5 10 125 3,375 1,81 0,5363 10 30 5 4 2 22 10 27
3039 isotropic with a soft zone 52 12,5 10 125 3,375 1,42 0,4207 10 60 5 4 2 22 10 27
3040 isotropic with a soft zone 52 12,5 10 125 3,375 1,5 0,4444 10 90 5 4 2 22 10 27
3041 isotropic with a hard zone 52 12,5 10 125 3,375 1,94 0,5748 10 0 5 22 10 4 2 27
3042 isotropic with a hard zone 52 12,5 10 125 3,375 1,47 0,4356 10 30 5 22 10 4 2 27
3043 isotropic with a hard zone 52 12,5 10 125 3,375 1,37 0,4059 10 60 5 22 10 4 2 27
3044 isotropic with a hard zone 52 12,5 10 125 3,375 1,33 0,3941 10 90 5 22 10 4 2 27
3045 isotropic with a variation in zone elastic parameters 52 12,5 10 125 1,08 0,001 0,0009 10 90 5 4 2 22 10 27
3046 isotropic with a variation in zone elastic parameters 52 12,5 10 125 1,08 0,89 0,8241 10 90 5 4 2 10 5 27
3047 isotropic with a variation in zone elastic parameters 52 12,5 10 125 1,08 1,41 1,3056 10 90 5 4 2 7 3 27
3048 isotropic with a variation in zone elastic parameters 52 12,5 10 125 1,08 1,72 1,5926 10 90 5 4 2 5 2,5 27
3049 isotropic with a variation in zone elastic parameters 52 12,5 10 125 1,08 1,88 1,7407 10 90 5 9 4 9 4 27
3050 isotropic with a variation in zone elastic parameters 52 12,5 10 125 1,08 1,66 1,537 10 90 5 10 5 22 10 27
3051 isotropic with a variation in zone elastic parameters 52 12,5 10 125 1,08 1,57 1,4537 10 90 5 7 3 22 10 27
3052 isotropic with a variation in zone elastic parameters 52 12,5 10 125 1,08 1,5 1,3889 10 90 5 4 2 22 10 27
3053 isotropic with a variation in zone elastic parameters 52 12,5 10 125 1,08 1,42 1,3148 10 90 5 2 1 22 10 27
VALLEY
4031 isotropic with a soft zone 52 2,5 10 25 0,675 1,3 1,9259 10 0 5 4 2 22 10 27
4032 isotropic with a soft zone 52 2,5 10 25 0,675 1,39 2,0593 10 30 5 4 2 22 10 27
4033 isotropic with a soft zone 52 2,5 10 25 0,675 1,09 1,6148 10 60 5 4 2 22 10 27
4034 isotropic with a soft zone 52 2,5 10 25 0,675 0,93 1,3778 10 90 5 4 2 22 10 27
4035 isotropic with a hard zone 52 2,5 10 25 0,675 1,56 2,3111 10 0 5 22 10 4 2 27
4036 isotropic with a hard zone 52 2,5 10 25 0,675 1,27 1,8815 10 30 5 22 10 4 2 27
4037 isotropic with a hard zone 52 2,5 10 25 0,675 0 0 10 60 5 22 10 4 2 27
4038 isotropic with a hard zone 52 2,5 10 25 0,675 1,79 2,6519 10 90 5 22 10 4 2 27
4039 isotropic with a variation in zone elastic parameters 52 2,5 10 25 0,675 0,001 0,0015 10 90 5 4 2 22 10 27
4040 isotropic with a variation in zone elastic parameters 52 2,5 10 25 0,675 0,32 0,4741 10 90 5 4 2 10 5 27
4041 isotropic with a variation in zone elastic parameters 52 2,5 10 25 0,675 0,93 1,3778 10 90 5 4 2 7 3 27
4042 isotropic with a variation in zone elastic parameters 52 2,5 10 25 0,675 1,34 1,9852 10 90 5 4 2 5 2,5 27
4043 isotropic with a variation in zone elastic parameters 52 2,5 10 25 0,675 1,5 2,2222 10 90 5 9 4 9 4 27
4044 isotropic with a variation in zone elastic parameters 52 2,5 10 25 0,675 1,06 1,5704 10 90 5 10 5 22 10 27
4045 isotropic with a variation in zone elastic parameters 52 2,5 10 25 0,675 0,97 1,437 10 90 5 7 3 22 10 27
4046 isotropic with a variation in zone elastic parameters 52 2,5 10 25 0,675 0,93 1,3778 10 90 5 4 2 22 10 27
4047 isotropic with a variation in zone elastic parameters 52 2,5 10 25 0,675 0,89 1,3185 10 90 5 2 1 22 10 27
4048 isotropic with a soft zone and buried valley 52 2,5 10 25 0,675 1,3 1,9259 10 90 5 4 2 22 10 27 20
4049 isotropic with a hard zone and buried valley 52 2,5 10 25 0,675 0,9 1,3333 10 90 5 22 10 4 2 27 20
FLOW MODELS
Relative Flow difference Rock mass properties
Case no. Model Angle Ratio DT C σcmin Flow between coupled Deformation Bulk Shear Weight Permeability Porosity
of slope C/DT and uncoupled Modulus Modulus Modulus Rock initial initial
(deg.) (m) (m) (MPa) (%) (GPa) (GPa) (GPa) (kN/m3) (m2) (%)
FLAT
1100 coupled 0 3,7 6 22 0,594 1,76E-06 175,6 0,5 0,4 0,2 27 1,00E-12 0,001
1101 coupled 0 3,7 6 22 0,594 1,12E-06 74,4 1 0,8 0,4 27 1,00E-12 0,001
1102 coupled 0 3,7 6 22 0,594 8,25E-07 28,9 2 1,7 0,8 27 1,00E-12 0,001
1103 coupled 0 3,7 6 22 0,594 7E-07 9,4 5 4 2 27 1,00E-12 0,001
1104 coupled 0 3,7 6 22 0,594 6,56E-07 2,5 15 12,5 5,8 27 1,00E-12 0,001
1105 coupled 0 3,7 6 22 0,594 6,45E-07 0,7 32 27 12 27 1,00E-12 0,001
1106 uncoupled 0 3,7 6 22 0,594 6,4E-07 0 32 27 12 27 1,00E-12 0,001
PLASTIC MODELS WITH STRUCTURAL FEATURES
(DT=10m)
Structural feature Rock mass
Ratio C
Case Model Angle σcmin σmmin CAR (Thickness=10m; Distance from tunnel center=5m) Weight Elastic
no. of slope C/DT Angle Bulk Shear Friction Cohesion Dilation Tension Bulk Shear Friction Cohesion Dilation Tension Rock model
(deg.) (m) (MPa) (MPa) (deg) (GPa) (GPa) (Deg.) (kPa) (Deg.) (kPa) (GPa) (GPa) (Deg.) (kPa) (Deg.) (kPa) (kN/m3) case
FLAT
1200 isotropic with a soft zone 0 4,5 45 1,215 0,1 0,08 90 4 2 25 25 8 25 22 10 40 1000 25 1000 27 1047
1201 isotropic with a hard zone 0 4,5 45 1,215 0,1 0,08 90 22 10 40 1000 25 1000 4 2 25 25 8 25 27 1059
SLOPE
2100 isotropic with a soft zone 45 4 40 0,706 0 0 135 4 2 25 25 8 25 22 10 40 1000 25 1000 27 2048
2101 isotropic with a hard zone 45 4 40 0,706 0 0 135 22 10 40 1000 25 1000 4 2 25 25 8 25 27 2054
HILL
3100 isotropic with a soft zone 52 12,5 125 3,375 0,2 0,16 60 4 2 25 25 8 25 22 10 40 1000 25 1000 27 3039
3101 isotropic with a hard zone 52 12,5 125 3,375 0,6 0,49 60 22 10 40 1000 25 1000 4 2 25 25 8 25 27 3043
VALLEY
4100 isotropic with a soft zone 52 2,5 25 0,675 0 0 60 4 2 25 25 8 25 22 10 40 1000 25 1000 27 4033
4101 isotropic with a hard zone 52 2,5 25 0,675 0,5 0,74 60 22 10 40 1000 25 1000 4 2 25 25 8 25 27 4037
APPENDIX B
FLAC numerical codes
B.1 Tunnel geometry

B1.1 D – shaped

config
grid 90,90
gen (-30.0,-30.0) (-30.0,30.0) (30.0,30.0) (30.0,-30.0) i 21 71 j 21 71
gen (-30.0,30.0) (-100.0,100.0) (100.0,100.0) (30.0,30.0) i 21 71 j 71 91
gen (-100.0,-100.0) (-100.0,100.0) (-30.0,30.0) (-30.0,-30.0) i 1 21 j 21 71
gen (-100.0,-100.0) (-30.0,-30.0) (30.0,-30.0) (100.0,-100.0) i 21 71 j 1 21
gen (30.0,-30.0) (30.0,30.0) (100.0,100.0) (100.0,-100.0) i 71 91 j 21 71
model elastic i=21,70 j=21,70
model elastic i=21,70 j=71,90
model elastic i=1,20 j=21,70
model elastic i=21,70 j=1,20
model elastic i=71,90 j=21,70
attach aside from 21 1 to 21 21 bside from 1 21 to 21 21
attach aside from 21 91 to 21 71 bside from 1 71 to 21 71
attach aside from 71 1 to 71 21 bside from 91 21 to 71 21
attach aside from 71 91 to 71 71 bside from 91 71 to 71 71

prop density=2700.0 bulk=1.E9 shear=3.0E8


table 1 -3.84,-5.77 -3.85,3.09 -2.67,4.20 -1.2,4.86
table 1 0,5.00 1.2,4.86 2.67,4.20 3.85,3.09
table 1 3.84,-5.77 -3.84,-5.77

gen t 1

fix y i 1 90 j 1
fix x i 1 j 1 90
fix x i 90 j 1 90
set grav=9.81
initial sxx -5.4e6 var=0,5.4e6
model null reg 45 46

solve

--------------------------------------------------------
B1.2 – Circle tunnel

config
grid 80,80
gen (-30.0,-30.0) (-30.0,30.0) (30.0,30.0) (30.0,-30.0) i 21 61 j 21 61
gen (-30.0,30.0) (-100.0,100.0) (100.0,100.0) (30.0,30.0) i 21 61 j 61 81
gen (-100.0,-100.0) (-100.0,100.0) (-30.0,30.0) (-30.0,-30.0) i 1 21 j 21 61
gen (-100.0,-100.0) (-30.0,-30.0) (30.0,-30.0) (100.0,-100.0) i 21 61 j 1 21
gen (30.0,-30.0) (30.0,30.0) (100.0,100.0) (100.0,-100.0) i 61 81 j 21 61
model elastic i=21,60 j=21,60
model elastic i=21,60 j=61,80
model elastic i=1,20 j=21,60
model elastic i=21,60 j=1,20
model elastic i=61,80 j=21,60
attach aside from 21 1 to 21 21 bside from 1 21 to 21 21
attach aside from 21 81 to 21 61 bside from 1 61 to 21 61
attach aside from 61 1 to 61 21 bside from 81 21 to 61 21
attach aside from 61 81 to 61 61 bside from 81 61 to 61 61
prop density=2700.0 bulk=1.E9 shear=3.0E8
gen circle 0 0 5
fix y i 1 80 j 1
fix x i 1 j 1 80
fix x i 80 j 1 80
set grav=9.81
initial sxx -5.4e6 var=0,5.4e6
model null reg 40 41

solve

--------------------------------------------------------
B2 – Isotropic models

B2.1 - Flat

config
grid 80,80
gen (-30.0,-30.0) (-30.0,30.0) (30.0,30.0) (30.0,-30.0) i 21 61 j 21 61
gen (-30.0,30.0) (-100.0,100.0) (100.0,100.0) (30.0,30.0) i 21 61 j 61 81
gen (-100.0,-100.0) (-100.0,100.0) (-30.0,30.0) (-30.0,-30.0) i 1 21 j 21 61
gen (-100.0,-100.0) (-30.0,-30.0) (30.0,-30.0) (100.0,-100.0) i 21 61 j 1 21
gen (30.0,-30.0) (30.0,30.0) (100.0,100.0) (100.0,-100.0) i 61 81 j 21 61
model elastic i=21,60 j=21,60
model elastic i=21,60 j=61,80
model elastic i=1,20 j=21,60
model elastic i=21,60 j=1,20
model elastic i=61,80 j=21,60
attach aside from 21 1 to 21 21 bside from 1 21 to 21 21
attach aside from 21 81 to 21 61 bside from 1 61 to 21 61
attach aside from 61 1 to 61 21 bside from 81 21 to 61 21
attach aside from 61 81 to 61 61 bside from 81 61 to 61 61
prop density=2700.0 bulk=1.E9 shear=3.0E8
gen circle 0 0 5
fix y i 1 80 j 1
fix x i 1 j 1 80
fix x i 80 j 1 80
set grav=9.81
initial sxx -5.4e6 var=0,5.4e6
model null reg 40 41

solve

--------------------------------------------------------
B2.2 - Slope

config
grid 100,150
model elastic
prop density=2700.0 bulk=1.E9 shear=3.0E8
gen 0,0 0,50 100,50 100,0 j 1,51
gen line 0.0,150.0 100.0,50.0
gen circle 50,40 5
fix y i 1 100 j 1
fix x i 1 j 1 150
fix x i 100 j 1 150
set grav=9.81
initial sxx -4.05e6 var=0,4.05e6

solve

mod null reg 100,150

solve

mod null reg 50,40


solve
save C:\ajr\phd\flac\tcs\slope\sl45\K1\sl45-K1-5.sav

--------------------------------------------------------
B2.3 - Hill

new
config
grid 200,160
model elastic
prop den=2700 bulk=1e9 shear=.3e9
gen -100,-50 -100,110 100,110 100,-50
gen line -100,20.0 -80,28
gen line -80,28 -60,41
gen line -60,41 -40,61
gen line -40,61 -20,86
gen line -20,86 -10,96
gen line -10,96 0,100.0
gen line 0,100.0 10,96
gen line 10,96 20,86.0
gen line 20,86.0 40,61
gen line 40,61 60,41
gen line 60,41 80,28
gen line 80,28 100,20.0

gen circle 0,-30 5

fix y i 1 200 j 1
fix x i 1 j 1 160
fix x i 201 j 1 160
set grav=9.81
initial sxx -4.32e6 var=0,4.32e6
initial syy -4.32e6 var=0,4.32e6
solve

mod null reg 1,159

solve

model null reg 100,20

solve
step 100

--------------------------------------------------------
B2.4 - Valley

new
config
grid 200,160
model elastic
prop den=2700 bulk=1e9 shear=.3e9
gen -100,-50 -100,110 100,110 100,-50
gen line -100,80.0 -80,72
gen line -80,72 -60,59
gen line -60,59 -40,39
gen line -40,39 -20,14
gen line -20,14 -10,4
gen line -10,4 0,0.0
gen line 0,0.0 10,4
gen line 10,4 20,14.0
gen line 20,14.0 40,39
gen line 40,39 60,59
gen line 60,59 80,72
gen line 80,72 100,80.0

gen circle 0,-30 5

fix y i 1 200 j 1
fix x i 1 j 1 160
fix x i 201 j 1 160
set grav=9.81

initial syy -4.32e6 var=0,4.32e6


initial sxx -4.32e6 var=0,4.32e6
solve
mod null reg 1,159

solve

model null reg 100,20

solve

--------------------------------------------------------
B2.5 – Buried Valley

new
config ext 1
grid 200,160
model elastic
prop den=2700 bulk=1e9 shear=.3e9
gen -100,-50 -100,110 100,110 100,-50
gen line -100,80.0 -80,72
gen line -80,72 -60,59
gen line -60,59 -40,39
gen line -40,39 -20,14
gen line -20,14 -10,4
gen line -10,4 0,0.0
gen line 0,0.0 10,4
gen line 10,4 20,14.0
gen line 20,14.0 40,39
gen line 40,39 60,59
gen line 60,59 80,72
gen line 80,72 100,80.0
gen line -80,72 80,72

gen circle 0,-30 5

fix y i 1 200 j 1
fix x i 1 j 1 160
fix x i 201 j 1 160
set grav=9.81

initial syy -4.32e6 var=0,4.32e6


initial sxx -4.32e6 var=0,4.32e6

solve

mod null reg 1,159

solve

model null reg 100,20

solve

mod elastic reg 100,70


prop den=2000 bulk=1e6 shear=.3e6 reg 100,70
solve

def k
loop i (1,izones)
loop j (1,jzones)
if model(i,j) # 1
_sx = sxx(i,j)
_sy = syy(i,j)
if _sy # 0.
ex_1(i,j) = _sx/_sy
else
ex_1(i,j) = 0. ; or some other value to indicate zero syy
endif
endif
endloop
endloop

end
k
B3 – Anisotropic models

B3.1 - Flat

new

;1 - INITIAL GRAVITY ; NO EXCAVATION

grid 100,100
model anisotropic

set mech on
gen circle 50 50 5

prop density=2.7e-3 shear=1.2E3


prop angle=60 xm=10.7e3 ym=5.2e3 nuy=0.2 nuz=0.41
fix y i 1 100 j 1
fix x i 1 j 1 100
fix x i 101 j 1 100
set grav=9.81
initial sxx -2.7 var=0,2.7
initial syy -2.7 var=0,2.7

solve

;3 - EXCAVATION

model null reg 50 51


model null reg 50 54

solve

--------------------------------------------------------
B3.2 - Slope

new

;1 - INITIAL GRAVITY ; NO EXCAVATION


grid 100,150
model anisotropic

set mech on
gen circle 50 40 5

gen line 0,150 100,50

prop density=2.7e-3 shear=1.2E3


prop angle=60 xm=10.e3 ym=2e3 nuy=0.2 nuz=0.41
fix y i 1 100 j 1
fix x i 1 j 1 150
fix x i 101 j 1 150
set grav=9.81
initial sxx -4.05 var=0,4.05
initial syy -4.05 var=0,4.05

solve

;2 - EROSION

mod null reg 20,149

solve

;3 - EXCAVATION

model null reg 50 41

solve

--------------------------------------------------------
B3.3 - Hill

new

;1 - INITIAL GRAVITY ; NO EXCAVATION

grid 100,80
model anisotropic
set mech on

gen -50,-25 -50,55 50,55 50,-25


gen line -50,10.0 -40,14
gen line -40,14 -30,20
gen line -30,20 -20,30
gen line -20,30 -10,43
gen line -10,43 -5,48
gen line -5,48 0,50.0
gen line 0,50.0 5,48
gen line 5,48 10,43
gen line 10,43 20,30
gen line 20,30.0 30,20
gen line 30,20 40,14
gen line 40,14 50,10

gen circle 0 -15 3

prop density=2.7e-3 shear=1.2E3


prop angle=60 xm=10.e3 ym=2e3 nuy=0.2 nuz=0.41
fix y i 1 100 j 1
fix x i 1 j 1 80
fix x i 101 j 1 80
set grav=9.81
initial sxx -2.16 var=0,2.16
initial syy -2.16 var=0,2.16
solve

;2 - EROSION

mod null reg 2,78


solve

;3 - EXCAVATION

model null reg 50 11


solve
--------------------------------------------------------
B3.4 - Valley

new

;1 - INITIAL GRAVITY ; NO EXCAVATION

grid 100,80
model anisotropic
set mech on

gen -50,-25 -50,55 50,55 50,-25


gen line -50,40.0 -40,36
gen line -40,36 -30,30
gen line -30,30 -20,20
gen line -20,20 -10,7
gen line -10,7 -5,2
gen line -5,2 0,0.0
gen line 0,0.0 5,2
gen line 5,2 10,7.0
gen line 10,7.0 20,20
gen line 20,20 30,30
gen line 30,30 40,36
gen line 40,36 50,40.0

gen circle 0 -15 3

prop density=2.7e-3 shear=1.2E3


prop angle=60 xm=10.e3 ym=2e3 nuy=0.2 nuz=0.41
fix y i 1 100 j 1
fix x i 1 j 1 80
fix x i 101 j 1 80
set grav=9.81
initial sxx -2.16 var=0,2.16
initial syy -2.16 var=0,2.16
solve

;2 - EROSION

mod null reg 2,78


solve

;3 - EXCAVATION

model null reg 50 11


solve
--------------------------------------------------------
B4 – Structural features

B4.1 - Flat

new

;1 - INITIAL GRAVITY ; NO EXCAVATION

grid 100,100
model elastic
set mech on

prop density=2700.0 bulk=22E9 shear=10E9

gen line 0,84 100,26


gen line 0,87 100,29

prop density=2700.0 bulk=4e9 shear=2e9 reg 96 30

gen circle 50 50 5

fix y i 1 100 j 1
fix x i 1 j 1 100
fix x i 101 j 1 100
set grav=9.81
initial sxx -2.7e6 var=0,2.7e6
initial syy -2.7e6 var=0,2.7e6

solve

;3 - EXCAVATION

model null reg 50 51

solve

--------------------------------------------------------
B4.2 - Slope

new
config
grid 100,150
model elastic
prop density=2700.0 bulk=22E9 shear=10E9
gen 0,0 0,50 100,50 100,0 j 1,51

gen line 0.0,150.0 8,142


gen line 8,142 100.0,50.0

gen line 0,135 78,0


gen line 4,150 8,142
gen line 8,142 89,0

gen circle 50,40 5

fix y i 1 100 j 1
fix x i 1 j 1 150
fix x i 100 j 1 150
set grav=9.81
initial sxx -4.05e6 var=0,4.05e6
initial syy -4.05e6 var=0,4.05e6

prop bulk=4e9 shear=2e9 reg 75,14


prop bulk=4e9 shear=2e9 reg 3,149

solve

mod null reg 100,150


mod null reg 3,149

solve

mod null reg 50,40

solve

--------------------------------------------------------
B4.3 - Hill

new
config
grid 200,160
model elastic
prop den=2700 bulk=22e9 shear=10e9
gen -100,-50 -100,110 100,110 100,-50
gen line -75,110 -44,57
gen line -44,57 17,-50
gen line -63,110 -37,65
gen line -37,65 28,-50
gen line -100,20.0 -80,28
gen line -80,28 -60,41
gen line -60,41 -44,57
gen line -44,57 -37,65
gen line -37,65 -20,86
gen line -20,86 -10,96
gen line -10,96 0,100.0
gen line 0,100.0 10,96
gen line 10,96 20,86.0
gen line 20,86.0 40,61
gen line 40,61 60,41
gen line 60,41 80,28
gen line 80,28 100,20.0

prop bulk=4e9 shear=2e9 reg 44 140


prop bulk=4e9 shear=2e9 reg 116 11

gen circle 0,-30 5

fix y i 1 200 j 1
fix x i 1 j 1 160
fix x i 201 j 1 160
set grav=9.81
initial sxx -4.32e6 var=0,4.32e6
initial syy -4.32e6 var=0,4.32e6
solve

mod null reg 1,159


mod null reg 44,140
mod null reg 170,120
solve
model null reg 100,20
solve
--------------------------------------------------------
B4.4 - Valley

new
config
grid 200,160
model elastic
prop den=2700 bulk=22e9 shear=10e9
gen -100,-50 -100,110 100,110 100,-50
gen line -75,110 -19,13
gen line -19,13 17,-50
gen line -63,110 0,0
gen line 0,0 28,-50
gen line -100,80.0 -80,72
gen line -80,72 -60,59
gen line -60,59 -40,39
gen line -40,39 -20,14
gen line -20,14 -10,4
gen line -10,4 0,0.0
gen line 0,0.0 10,4
gen line 10,4 20,14.0
gen line 20,14.0 40,39
gen line 40,39 60,59
gen line 60,59 80,72
gen line 80,72 100,80.0

prop bulk=4e9 shear=2e9 reg 38 147


prop bulk=4e9 shear=2e9 reg 119 7

gen circle 0,-30 5

fix y i 1 200 j 1
fix x i 1 j 1 160
fix x i 201 j 1 160
set grav=9.81
initial sxx -4.32e6 var=0,4.32e6
initial syy -4.32e6 var=0,4.32e6
solve
mod null reg 1,159
mod null reg 36,152
mod null reg 195,157
solve
model null reg 100,20
solve
B4.5 – Buried Valley

new
config ext 1
grid 200,160
model elastic
prop den=2700 bulk=22e9 shear=10e9
gen -100,-50 -100,110 100,110 100,-50

gen line 4,110 4,2


gen line 4,2 4,-50
gen line 14,110 14,8
gen line 14,8 14,-50

gen line -100,80.0 -80,72


gen line -80,72 -60,59
gen line -60,59 -40,39
gen line -40,39 -20,14
gen line -20,14 -10,4
gen line -10,4 0,0.0
gen line 0,0.0 10,4
gen line 10,4 20,14.0
gen line 20,14.0 40,39
gen line 40,39 60,59
gen line 60,59 80,72
gen line 80,72 100,80.0

gen line -80,72 80,72

prop bulk=4e9 shear=2e9 reg 110,50

gen circle 0,-30 5

fix y i 1 200 j 1
fix x i 1 j 1 160
fix x i 201 j 1 160
set grav=9.81
initial sxx -4.32e6 var=0,4.32e6
initial syy -4.32e6 var=0,4.32e6
solve

mod null reg 1,159


mod null reg 108,155
mod null reg 195,157

solve

model null reg 100,20

solve

mod elastic reg 70,105


mod elastic reg 108,105
mod elastic reg 120,105

prop den=2000 bulk=1e6 shear=.3e6 reg 70,105


prop den=2000 bulk=1e6 shear=.3e6 reg 108,105
prop den=2000 bulk=1e6 shear=.3e6 reg 120,105

solve

def k
loop i (1,izones)
loop j (1,jzones)
if model(i,j) # 1
_sx = sxx(i,j)
_sy = syy(i,j)
if _sy # 0.
ex_1(i,j) = _sx/_sy
else
ex_1(i,j) = 0. ; or some other value to indicate zero syy
endif
endif
endloop
endloop

end
k
B5.1 – Coupled flow models – Flat topography

new
config gw

def istrain

gam=27 ;kN/m3
h=25 ;m center of tunnel elevation
Erm=1e6 ;kPa
inistr=-(gam*h)/Erm
end
istrain

def tabn
mn=5.2
strain=inistr+strain_f
loop n (1,2001)
xtable(1,n)=strain
ytable(1,n)=start_n*exp(-(mn*(strain-inistr))/strain_f)
strain=strain+0.00001
end_loop
end
set start_n=0.001 strain_f=-0.01
tabn

def tabk
mk=10
strain=inistr+strain_f
loop n (1,2001)
xtable(2,n)=strain
ytable(2,n)=start_k*(1-((strain-inistr)/strain_f))^mk
strain=strain+0.00001
end_loop
end
set start_k=1e-12 strain_f=-0.01
tabk

;1 - INITIAL GRAVITY ; NO EXCAVATION

grid 50,50
model elastic

set flow off


set mech on

gen circle 25 25 3

prop density=2700.0 bulk=0.8E9 shear=0.4E9


fix y i 1 50 j 1
fix x i 1 j 1 50
fix x i 51 j 1 50
set grav=9.81
initial sxx -1.35e6 var=0,1.35e6

solve

;2 - INITIAL FLUID ; NO EXCAVATION

set flow on

prop por=start_n por_tab=1 perm=start_k per_tab=2


water bulk=1e6 dens 1000
ini pp 0.5e6 var 0 -0.5e6
fix pp i 1
fix sat i 1
fix pp i 50
fix sat i 50
fix pp j 50
free sat j 50
free pp j 1
free sat j 1

solve

;3 - EXCAVATION - NO FLUID

set flow off

model null reg 25 26

solve

;4 - EXCAVATION - FLUID DRAINAGE

set flow on
unmark

gen circle 25 25 3
fix pp mark
ini pp=1.25e6 mark

solve

def flow
flow1=gflow(25,28)+gflow(26,28)+gflow(27,28)+gflow(27,27)+gflow(28,27)
flow2=gflow(28,26)+gflow(28,25)+gflow(28,24)+gflow(28,23)+gflow(27,23)
flow3=gflow(27,22)+gflow(26,22)+gflow(25,22)+gflow(24,22)+gflow(23,22)
flow4=gflow(23,23)+gflow(22,23)+gflow(22,24)+gflow(22,25)+gflow(22,26)
flow5=gflow(22,27)+gflow(23,27)+gflow(23,28)+gflow(24,28)
flow=flow1+flow2+flow3+flow4+flow5
end
hist flow
print flow
print tabk
;---------------------------------------------------------------------------
B5.2 – Uncoupled flow models – Flat topography

new
config gw

;1 - INITIAL GRAVITY ; NO EXCAVATION

grid 50,50
model elastic

set flow off


set mech on

gen circle 25 25 3

prop density=2700.0 bulk=27E9 shear=12E9


fix y i 1 50 j 1
fix x i 1 j 1 50
fix x i 51 j 1 50
set grav=9.81
initial sxx -1.35e6 var=0,1.35e6

solve

;2 - INITIAL FLUID ; NO EXCAVATION

set flow on

prop por=0.001 perm=1e-12


water bulk=1e6 dens 1000
ini pp 0.5e6 var 0 -0.5e6
fix pp i 1
fix sat i 1
fix pp i 50
fix sat i 50
fix pp j 50
free sat j 50
free pp j 1
free sat j 1
solve

;3 - EXCAVATION - NO FLUID

set flow off

model null reg 25 26

solve

;4 - EXCAVATION - FLUID DRAINAGE

set flow on

unmark

gen circle 25 25 3
fix pp mark
ini pp=1.25e6 mark

solve

def flow
flow1=gflow(25,28)+gflow(26,28)+gflow(27,28)+gflow(27,27)+gflow(28,27)
flow2=gflow(28,26)+gflow(28,25)+gflow(28,24)+gflow(28,23)+gflow(27,23)
flow3=gflow(27,22)+gflow(26,22)+gflow(25,22)+gflow(24,22)+gflow(23,22)
flow4=gflow(23,23)+gflow(22,23)+gflow(22,24)+gflow(22,25)+gflow(22,26)
flow5=gflow(22,27)+gflow(23,27)+gflow(23,28)+gflow(24,28)
flow=flow1+flow2+flow3+flow4+flow5
end
hist flow
print flow

;---------------------------------------------------------------------
B6 – Plastic models

B6.1 – Flat
new

;1 - INITIAL GRAVITY ; NO EXCAVATION

grid 100,100
model m

set mech on

gen circle 50 50 5

gen line 54,0 54,100


gen line 64,0 64,100

prop density=2700.0 bulk=22E9 shear=10E9 fric 40 co 1e6 dil 25 ten 1e6


fix y i 1 100 j 1
fix x i 1 j 1 100
fix x i 101 j 1 100
set grav=9.81
initial sxx -2.7e6 var=0,2.7e6
initial syy -2.7e6 var=0,2.7e6
initial szz -2.7e6 var=0,2.7e6

prop density=2700.0 bulk=4e9 shear=2e9 fric 25 co 25e3 dil 8 ten 25e3 i=54,64

solve step 5000

;3 - EXCAVATION

model null reg 50 51


model null reg 54 50

solve step 10000


B6.2 – Slope
new
config
grid 100,150
model m
prop density=2700.0 bulk=22E9 shear=10E9 fric 40 co 1e6 dil 25 ten 1e6
gen 0,0 0,50 100,50 100,0 j 1,51

gen line 0,150 68,82


gen line 68,82 79,71
gen line 79,71 100,50

gen line 0,10 68,78


gen line 68,78 68,82
gen line 68,82 71,82
gen line 71,82 100,110

gen line 4,0 75,72


gen line 75,72 79,71
gen line 79,71 79,75
gen line 79,75 100,96

gen circle 50,40 5

fix y i 1 100 j 1
fix x i 1 j 1 150
fix x i 100 j 1 150
set grav=9.81
initial sxx -4.05e6 var=0,4.05e6
initial syy -4.05e6 var=0,4.05e6
initial szz -4.05e6 var=0,4.05e6

prop bulk=4e9 shear=2e9 fric 25 co 25e3 dil 8 ten 25e3 reg 20,24
prop bulk=4e9 shear=2e9 fric 25 co 25e3 dil 8 ten 25e3 reg 90,92

solve step 5000

mod null reg 20,149


mod null reg 92,96
mod null reg 90,73

solve step 5000

mod null reg 50,40

solve step 10000


B6.3 – Hill
new
config
grid 200,160
model m
prop den=2700 bulk=22e9 shear=10e9 fric 40 co 1e6 dil 25 ten 1e6
gen -100,-50 -100,110 100,110 100,-50

gen line -75,110 -44,57


gen line -44,57 17,-50
gen line -63,110 -37,65
gen line -37,65 28,-50
gen line -100,20.0 -80,28
gen line -80,28 -60,41
gen line -60,41 -44,57
gen line -44,57 -37,65
gen line -37,65 -20,86
gen line -20,86 -10,96
gen line -10,96 0,100.0
gen line 0,100.0 10,96
gen line 10,96 20,86.0
gen line 20,86.0 40,61
gen line 40,61 60,41
gen line 60,41 80,28
gen line 80,28 100,20.0

prop bulk=4e9 shear=2e9 fric 25 co 25e3 dil 8 ten 25e3 reg 44 140
prop bulk=4e9 shear=2e9 fric 25 co 25e3 dil 8 ten 25e3 reg 116 11

gen circle 0,-30 5

fix y i 1 200 j 1
fix x i 1 j 1 160
fix x i 201 j 1 160
set grav=9.81
initial sxx -4.32e6 var=0,4.32e6
initial syy -4.32e6 var=0,4.32e6
initial szz -4.32e6 var=0,4.32e6
solve step 5000

mod null reg 1,159


mod null reg 44,140
mod null reg 170,120
solve step 5000
model null reg 100,20
solve step 10000
B6.4 – Valley
new
config
grid 200,160
model m
prop den=2700 bulk=22e9 shear=10e9 fric 40 co 1e6 dil 25 ten 1e6
gen -100,-50 -100,110 100,110 100,-50

gen line -75,110 -19,13


gen line -19,13 17,-50
gen line -63,110 0,0
gen line 0,0 28,-50
gen line -100,80.0 -80,72
gen line -80,72 -60,59
gen line -60,59 -40,39
gen line -40,39 -20,14
gen line -20,14 -10,4
gen line -10,4 0,0.0
gen line 0,0.0 10,4
gen line 10,4 20,14.0
gen line 20,14.0 40,39
gen line 40,39 60,59
gen line 60,59 80,72
gen line 80,72 100,80.0

prop bulk=4e9 shear=2e9 fric 25 co 25e3 dil 8 ten 25e3 reg 38 147
prop bulk=4e9 shear=2e9 fric 25 co 25e3 dil 8 ten 25e3 reg 119 7

gen circle 0,-30 5

fix y i 1 200 j 1
fix x i 1 j 1 160
fix x i 201 j 1 160
set grav=9.81
initial sxx -4.32e6 var=0,4.32e6
initial syy -4.32e6 var=0,4.32e6
initial szz -4.32e6 var=0,4.32e6
solve step 5000

mod null reg 1,159


mod null reg 36,152
mod null reg 195,157
solve step 5000
model null reg 100,20
solve step 10000
APPENDIX C
Models basic geometry and meshing
Figure C.1 – Circular tunnel

Figure C.2 – D-Shaped tunnel


Figure C.3 – Isotropic models with flat topography

Figure C.4 – Anisotropic models with flat topography


Figure C.5 - Isotropic and anisotropic models with slope topography ( β = 5° )

Figure C.6 - Isotropic and anisotropic models with slope topography ( β = 30° )
Figure C.7 - Isotropic and anisotropic models with slope topography ( β = 45° )
Figure C.8 - Isotropic and anisotropic models with hill topography ( β = 64° )

Figure C.9 - Isotropic and anisotropic models with hill topography ( β = 52° )
Figure C.10 - Isotropic and anisotropic models with hill topography ( β = 42° )

Figure C.11 - Isotropic and anisotropic models with hill topography ( β = 32° )
Figure C.12 - Isotropic and anisotropic models with hill topography ( β = 25° )

Figure C.13 - Isotropic and anisotropic models with hill topography ( β = 17° )
Figure C.14 - Isotropic and anisotropic models with valley topography ( β = 64° )

Figure C.15 - Isotropic and anisotropic models with valley topography ( β = 52° )
Figure C.16 - Isotropic and anisotropic models with valley topography ( β = 42° )

Figure C.17 - Isotropic and anisotropic models with valley topography ( β = 32° )
Figure C.18 - Isotropic and anisotropic models with valley topography ( β = 25° )

Figure C.19 - Isotropic and anisotropic models with valley topography ( β = 17° )
Figure C.20 - Isotropic model with flat topography and with tabular structural features
(D = 0; thickness = 3 m)

Figure C.21 - Isotropic model with flat topography and with tabular structural features
(D = 3; thickness = 3 m)

Figure C.22 - Isotropic model with flat topography and with tabular structural features
(D = 5; thickness = 3 m)
Figure C.23 - Isotropic model with flat topography and with tabular structural features
(D = 7; thickness = 3 m)

Figure C.24 - Isotropic model with flat topography and with tabular structural features
(D = 10; thickness = 3 m)

Figure C.25 - Isotropic model with flat topography and with tabular structural features
(D = 15; thickness = 3 m)
Figure C.26 - Isotropic model with flat topography and with tabular structural features
(D = 20; thickness = 3 m)

Figure C.27 - Isotropic model with flat topography and with tabular structural features
(D = 25; thickness = 3 m)

Figure C.28 - Isotropic model with flat topography and with tabular structural features
(D = 3; thickness = 2 m; 3 m)
Figure C.29 - Isotropic model with flat topography and with tabular structural features
(D = 3; thickness = 5 m; 10 m)

Figure C.30 - Isotropic model with flat topography and with tabular structural features
(D = 3; thickness = 15 m; 20 m)

Figure C.31 - Isotropic model with flat topography and with tabular structural features
(D = 3; thickness = 25 m; inf)
Figure C.32 - Isotropic model with flat topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 0°)

Figure C.33 - Isotropic model with flat topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 30°)
Figure C.34 - Isotropic model with flat topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 60°)

Figure C.35 - Isotropic model with flat topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 90°)
Figure C.36 - Isotropic model with slope topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 0°)

Figure C.37 - Isotropic model with slope topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 30°)
Figure C.38 - Isotropic model with slope topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 60°)

Figure C.39 - Isotropic model with slope topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 90°)
Figure C.40 - Isotropic model with hill topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 0°)

Figure C.41 - Isotropic model with hill topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 30°)
Figure C.42 - Isotropic model with hill topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 60°)

Figure C.43 - Isotropic model with hill topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 90°)
Figure C.44 - Isotropic model with valley topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 0°)

Figure C.45 - Isotropic model with valley topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 30°)
Figure C.46 - Isotropic model with valley topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 60°)

Figure C.47 - Isotropic model with valley topography and with tabular structural features
(D = 5; thickness = 10 m; orientation δ = 90°)
Figure C.48 - Isotropic fluid-mechanical interaction model with flat topography

Figure C.49 - Plastic model with flat topography and with vertical structural features
(D = 5; thickness = 10 m; orientation δ = 90°)
Figure C.50 - Plastic model with slope topography and with vertical structural features
(D = 5; thickness = 10 m; orientation δ = 135°)

Figure C.51 - Plastic model with hill topography and with vertical structural features
(D = 5; thickness = 10 m; orientation δ = 60°)
Figure C.52 - Plastic model with valley topography and with vertical structural features
(D = 5; thickness = 10 m; orientation δ = 60°)
APPENDIX D
Minimum stress (σmmin) contour lines showing stress perturbation
Figure D.1 – Circular models; cases 1 to 11 and cases 1001 to 1011

Figure D.2 – D shaped models; cases 12 to 22


Figure D.3 – Slope models (β = 5°); cases 2001 to 2008

Figure D.4 – Slope models (β = 30°); cases 2009 to 2016


Figure D.5 – Slope models (β = 45°); cases 2017 to 2025
Figure D.6 – Hill models (β = 64°); cases 3001 to 3005

Figure D.7 – Hill models (β = 52°); cases 3006 to 3010


Figure D.8 – Hill models (β = 42°); cases 3011 to 3015

Figure D.9 – Hill models (β = 32°); cases 3016 to 3020


Figure D.10 – Hill models (β = 25°); cases 3021 to 3025

Figure D.11 – Hill models (β = 17°); cases 3026 to 3030


Figure D.12 – Valley models (β = 64°); cases 4001 to 4004

Figure D.13 – Valley models (β = 52°); cases 4005 to 4008


Figure D.14 – Valley models (β = 42°); cases 4009 to 4012

Figure D.15 – Valley models (β = 32°); cases 4013 to 4016


Figure D.16 – Valley models (β = 25°); cases 4017 to 4020

Figure D.17 – Valley models (β = 17°); cases 4021 to 4024


a) D = 0 m b) D = 3 m

c) D = 5 m d) D = 7 m

e) D = 10 m f) D = 15 m

g) D = 20 m h) D = 25 m

Figure D.18 – Structural models, Soft vertical tabular zone (T = 3 m) with distance from
center variation; cases 1022 to 1030
a) D = 0 m b) D = 3 m

c) D = 5 m d) D = 7 m

e) D = 10 m f) D = 15 m

g) D = 20 m h) D = 25 m

Figure D.19 – Structural models, Soft horizontal tabular zone (T = 3 m) with distance
from center variation; cases 1031 to 1039
a) T = 2 m b) T = 3 m

c) T = 5 m d) T = 10 m

e) T = 15 m f) T = 20 m

g) T = 25 m h) T = inf. (contact)

Figure D.20 – Structural models, Soft vertical tabular zone (D = 3m) with thickness
variation; cases 1064 to 1071
a) δ = 0° b) δ = 30°

c) δ = 60° d) δ = 90°

Figure D.21 – Structural models, Soft tabular zone (T = 10 m) with angle variation; cases
1044 to 1047

a) δ = 0° b) δ = 30°

c) δ = 60° d) δ = 90°

Figure D.22 – Structural models, Hard tabular zone (T = 10 m) with angle variation;
cases 1056 to 1059
a) δ = 0° b) δ = 30°

c) δ = 60° d) δ = 90°

c) δ = 135°

Figure D.23 – Structural models, Soft tabular zone (T = 10 m) with angle variation and a
45° slope; cases 2044 to 2048
a) δ = 0° b) δ = 30°

c) δ = 60° d) δ = 90°

c) δ = 135°

Figure D.24 – Structural models, Hard tabular zone (T = 10 m) with angle variation and a
45° slope; cases 2050 to 2054
a) δ = 0° b) δ = 30°

c) δ = 60° d) δ = 90°

Figure D.25 – Structural models, Soft tabular zone (T = 10 m) with angle variation and a
52° hill; cases 3037 to 3040

a) δ = 0° b) δ = 30°

c) δ = 60° d) δ = 90°

Figure D.26 – Structural models, Hard tabular zone (T = 10 m) with angle variation and a
52° hill; cases 3041 to 3044
a) δ = 0° b) δ = 30°

c) δ = 60° d) δ = 90°

Figure D.27 – Structural models, Soft tabular zone (T = 10 m) with angle variation and a
52° valley; cases 4031 to 4034

a) δ = 0° b) δ = 30°

c) δ = 60° d) δ = 90°

Figure D.28 – Structural models, Hard tabular zone (T = 10 m) with angle variation and a
52° valley; cases 4035 to 4038
a) E1/E2 = 0.2 b) E1/E2 = 0.4

c) E1/E2 = 0.6 d) E1/E2 = 0.8

e) E1/E2 =2.2 f) E1/E2 = 3.3

g) E1/E2 =5 h) E1/E2 = 10

Figure D.29 – Structural models, Tabular zone (T = 10 m) with deformation modulus


variation for a flat model; cases 1072 to 1080
a) E1/E2 = 0.2 b) E1/E2 = 0.4

c) E1/E2 = 0.6 d) E1/E2 = 0.8

e) E1/E2 =2.2 f) E1/E2 = 3.3

g) E1/E2 =5 h) E1/E2 = 10

Figure D.30 – Structural models, Tabular zone (T = 10 m) with deformation modulus


variation for a slope model; cases 2055 to 2063
a) E1/E2 = 0.2 b) E1/E2 = 0.4

c) E1/E2 = 0.6 d) E1/E2 = 0.8

e) E1/E2 =2.2 f) E1/E2 = 3.3

g) E1/E2 =5 h) E1/E2 = 10

Figure D.31 – Structural models, Tabular zone (T = 10 m) with deformation modulus


variation for a hill model; cases 3045 to 3053
a) E1/E2 = 0.2 b) E1/E2 = 0.4

c) E1/E2 = 0.6 d) E1/E2 = 0.8

e) E1/E2 =2.2 f) E1/E2 = 3.3

g) E1/E2 =5 h) E1/E2 = 10

Figure D.32 – Structural models, Tabular zone (T = 10 m) with deformation modulus


variation for a valley model; cases 4039 to 4047
a) ω = 0° b) ω = 30°

a) ω = 60° b) ω = 90°

Figure D.33 – Structural models, Anisotropic with principal axis of anisotropy angle
variation for a flat model; cases 1012 to 1015
a) ω = 0° b) ω = 30°

c) ω = 60° d) ω = 90°

e) ω = 135°

Figure D.34 – Structural models, Anisotropic with principal axis of anisotropy angle
variation for a slope model; cases 2035 to 2039
a) ω = 0° b) ω = 30°

c) ω = 60° d) ω = 90°

Figure D.35 – Structural models, Anisotropic with principal axis of anisotropy angle
variation for a hill model; cases 3031 to 3034

a) ω = 0° b) ω = 30°

c) ω = 60° d) ω = 90°

Figure D.36 – Structural models, Anisotropic with principal axis of anisotropy angle
variation for a valley model; cases 3031 to 3034
a) Ex/Ey = 1 b) Ex/Ey = 1.3

c) Ex/Ey = 1.5 d) Ex/Ey = 1.8

e) Ex/Ey = 2 f) Ex/Ey = 4

Figure D.37 – Structural models, Anisotropic with variation of anisotropy intensity for a
flat model; cases 1016 to 1021
a) Pore pressure contours and flow vectors

b) Permeability, pore pressure, flow vector and volumetric strain

Figure D.38 – Flow models; cases 1016 to 1021


Figure D.39 – Plastic flat model, Soft vertical tabular zone (T = 10 m); cases 6001

Figure D.40 – Plastic slope model, Soft inclined tabular zone (T = 10 m); cases 6003
Figure D.41 – Plastic hill model, Soft inclined tabular zone (T = 10 m); cases 6005

Figure D.42 – Plastic valley model, Soft inclined tabular zone (T = 10 m); cases 6007
Figure D.43 – Plastic flat model, Hard vertical tabular zone (T = 10 m); cases 6002

Figure D.44 – Plastic slope model, Hard inclined tabular zone (T = 10 m); cases 6004
Figure D.45 – Plastic hill model, Hard inclined tabular zone (T = 10 m); cases 6006

Figure D.46 – Plastic valley model, Hard inclined tabular zone (T = 10 m); cases 6008

You might also like