Sie sind auf Seite 1von 91

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/328126754

Face stability of closed TBMs in Urban Tunnels

Thesis · December 2014

CITATIONS READS

2 94

1 author:

Senthilnath G T

18 PUBLICATIONS   12 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Thomson Line View project

All content following this page was uploaded by Senthilnath G T on 06 October 2018.

The user has requested enhancement of the downloaded file.


2nd level Specializing Master in
Tunnelling and Tunnel Boring Machines

Face Stability of Closed TBMs in Urban Tunnels

Senthilnath G T

Company supervisor Academic supervisor


Dr. Oskar Sigl – Geoconsult ZT GmbH Prof. Daniele Peila

Endorsed by

With the support of

December 2014
Face Stability of Closed TBMs in Urban
Tunnels

Senthilnath G T

2nd Level Specializing Masters in


Tunnelling and Tunnel Boring Machines
Politecnico di Torino, Italy

December 2014
Acknowledgements
This work was carried out during the on-the job training / Internship period from
July 2014 to November 2014 under the guidance of Prof. Daniele Peila with the help of
my employer Dr. Oskar Sigl.

I am very grateful to my supervisors who gave me the freedom and the trust to select
the research area myself and let me develop the things, which I considered important
from engineering practice point of view. The discussions, comments and the way of
communication of Dr. Oskar Sigl has been very motivating and inspiring.

I am indebted to International Tunnelling Association (ITA-AITES) and ITACET


Foundation, which sponsored this course work and provided the ITACET Scholarship
for my course work in Turin, Italy.

Special thanks to my colleagues in the master course for maintaining a comfortable


atmosphere and for making my time at Turin a wonderful experience.

Last, but not least, I would like to express my deep gratitude to my parents and parents-
in-law for their unconditional love. Of all people, I am most grateful to my wife, Sujitha
for all her patience, encouragement and love.

ii
Theory and calculation are not substitute for judgement, but are the basis for
sounder judgement.
- Ralph B. Peck

iii
Abstract
In recent years, several large size shield tunnels have been constructed in urban
areas at shallow depth with the increasing practical demands such as accessibility,
serviceability and economy. One of the main objectives of such tunnels is to adequately
support the soil and to minimize deformations during and after construction, because in
such conditions, tunnel collapse or extensive deformations can be catastrophic, and even
limited soil deformations may damage buildings. To prevent this it is necessary to
continuously support the excavated face. The analysis of face stability of any pressurized
shield machines are performed to avoid both, collapse (active failure) and blow-out
(passive failure) of the soil mass near the tunnel face. A number of studies have been
dedicated to tunnel face stability [1] and recent developments (like "multi-block
mechanism", "method of slices" etc.) have updated the widely used methods of analysis.
Here, in this study, around 30 face stability models (analytical, empirical and semi-
empirical) are studied and a comparison of minimum face pressure values computed
using recently developed methods are made.

A technique for performing Monte-Carlo simulation for probabilistic estimation of


minimum face pressure has been presented. The pseudo random number generator is
used to simulate failure times of individual units or modules in the system. Logical
expressions are then used to determine system success or failure. Failure probability
density function (PDF) and cumulative distribution function (CDF) of the overall
system are presented for different face stability models. Use of latest analytical models
with the above mentioned probabilistic techniques will result in realistic face pressure
values and thus avoid the operational disadvantages of over conservatively high face
pressure values like fluctuation of effective face pressure in chamber, high wear, arching
of muck at the entrance of screw conveyor [2]. Simplified model for implementing
probabilistic analysis with various face stability analytical models are explored and
presented in this study for quick implementation in design routines.

iv
Contents
Acknowledgements

Abstract

List of Figures

1 Introduction ........................................................................................... 1

1.1 Background ..................................................................................... 1

1.2 Objective and Scope of the Study ...................................................... 2

1.3 Organization of the report ................................................................ 2

2 Tunnel Face Stability .............................................................................. 3

2.1 Face Stability Models ....................................................................... 4

2.1.1 Models for Undrained Cohesive Soil (Tresca) ................................. 5

2.1.2 Models for Cohesionless Soil ........................................................ 13

2.1.3 Models for C-Phi Soil (Mohr Coulomb) ........................................ 20

2.2 Comparison of different models....................................................... 44

2.2.1 Cohesive Soils ............................................................................ 44

2.2.2 Cohesionless Soils ....................................................................... 46

2.2.3 C-Phi Soil .................................................................................. 49

2.2.4 Models which consider Seepage forces .......................................... 51

2.2.5 Models for Blowout Pressure ....................................................... 54

2.3 Case Study .................................................................................... 55

3 Probability Analysis .............................................................................. 59

v
3.1 Uncertainties ................................................................................. 61

3.2 Monte Carlo Simulation ................................................................. 62

3.3 Probability Assessment ................................................................... 63

4 Results and Conclusion ......................................................................... 72

Bibliography

Curriculum Vitae

Appendix A

vi
List of Figures
Figure 1. Failure mechanism observed in sand and clay [16].......................................... 5
Figure 2. Broms and Bennemark face stability model [17] ............................................. 6
Figure 3. Davis' face stability model (a) [11] .................................................................. 7
Figure 4. Davis' face stability model (b) [11] .................................................................. 7
Figure 5. Stability ratio design line proposed by Kumara & Mair [18] .......................... 8
Figure 6. Ellstein's face stability model [19] ................................................................... 9
Figure 7. Layout of plane strain heading problem [22] ................................................. 10
Figure 8. Failure mechanism proposed by [23].............................................................. 12
Figure 9. Stability number for 2D case [23] .................................................................. 13
Figure 10. Stability model of the method Atkinson and Potts [26] .............................. 14
Figure 11. Failure bulbs comparison with theoretical mechanism [27] ......................... 15
Figure 12. Improved 3D wedge prism model [28].......................................................... 16
Figure 13. Mechanical Equilibrium of the wedge [28] ................................................... 17
Figure 14. Distribution of the vertical stress σz along side slip surfaces [28] ................ 18
Figure 15. Design chart by Chen et al for cohesionless soils......................................... 19
Figure 16. Failure mechanism by Murayama [8]........................................................... 21
Figure 17. Statically admissible solution by Muehlhaus [12] ........................................ 22
Figure 18. Failure mechanics proposed by Krause [8]................................................... 22
Figure 19. Failure mechanics proposed by Mohkam [1] ................................................ 23
Figure 20. Conical blocks and kinematic conditions used in MI, MII and MIII [13] .... 24
Figure 21. 3D visualization of failure mechanism by Leca and Dormieux [37] ............. 24
Figure 22. Failure mechanics proposed by Leca and Dormieux [13] ............................. 25
Figure 23. Geometry of Mechanism MII [13] ................................................................ 26
Figure 24. Geometry of Mechanism MIII [13] ............................................................... 27
Figure 25. Stability model by Jancsecz & Steiner [33].................................................. 29
Figure 26. Seepage force f and effective support pressure s’ [2] .................................... 30
Figure 27. Stability model of Anagnostou & Kovari based on Horn [2] ....................... 30

vii
Figure 28. Nomograms for non-dimensional factors F0, F1 [2] ....................................... 31
Figure 29. Nomograms for non-dimensional factors F2, F3 [2] ...................................... 31
Figure 30. Stability model of Belter and Katzenbach [8] .............................................. 32
Figure 31. Multilayer wedge model by Broere [8] ......................................................... 32
Figure 32. Multilayer wedge model by Broere [8] ......................................................... 33
Figure 33. Extended Caquot's model [25] ..................................................................... 34
Figure 34. Excavation face as hemisphere [36].............................................................. 35
Figure 35. Multi-block failure mechanism [14] .............................................................. 36
Figure 36. Non-dimensional coefficients for face pressure estimation [14] .................... 37
Figure 37. Design chart for critical collapse pressure for C-Phi soil [14] ...................... 38
Figure 38. Cross section and horizontal section of a tunnel [9] ................................... 38
Figure 39. Force acting on infinitesimal slice [10] ......................................................... 39
Figure 40. Stability model used for analysis [10] .......................................................... 40
Figure 41. Design diagrams for effective support pressure, ∆h/H = 0-2 [10] .............. 43
Figure 42. Design diagrams for effective support pressure, ∆h/H = 3-5 [10] ................ 44
Figure 43. Comparison of face pressure for different cohesive soil model ..................... 45
Figure 44. Comparison of face pressure for different Cu values.................................... 46
Figure 45. Comparison of face pressure for Cohesionless soils ...................................... 47
Figure 46. Comparison of face pressure for different ϕ values ...................................... 48
Figure 47. Comparison of face pressure for c-ϕ soil ...................................................... 49
Figure 48. Comparison of face pressure for c-ϕ soil - influence of cohesion .................. 50
Figure 49. Comparison of effective face pressure as a function of ϕ’ ............................ 52
Figure 50. Comparison of effective face pressure as a function of c’ ............................. 53
Figure 51. Comparison of face pressure as a function of hydraulic gradient ................ 54
Figure 52. Comparison of Blowout pressure ................................................................. 55
Figure 53. Geological profile and tunnel alignment – Case study ................................ 56
Figure 54. Estimated TBM operating face pressure – Case study................................ 58
Figure 55. General stochastic problem [45] ................................................................... 61
Figure 56. COV of geotechnical parameters [45]........................................................... 61
Figure 57. Input values for MC simulation using Anagnostou, 2012 ............................ 64

viii
Figure 58. Snapshot of Random number generation ..................................................... 65
Figure 59. Histogram of MC simulation using Anagnostou, 2012................................. 66
Figure 60. CDF of MC simulation using Anagnostou, 2012 ......................................... 66
Figure 61. Summary of MC simulation using Anagnostou, 2012 in Excel .................... 67
Figure 62. Output of MC simulation using Chen et al 2014 model .............................. 68
Figure 63. Output of MC simulation using Mollon et al 2010 model ........................... 69
Figure 64. Distribution of input parameters for MC simulation ................................... 70
Figure 65. Comparison of PDF using the two methods ................................................ 70

ix
1 Introduction

1.1 Background

The problems posed by urban development and the challenges of traffic, utilities etc. are
often solved by digging tunnels. More and more tunnels are being excavated and there
are a host of new technologies to assist in excavation. Most urban tunnels are now shield
driven. This increases the importance of controlling the tunnel face pressure during its
advance. In the case of shallow tunnels, this pressure influences significantly the surface
settlements. This pressure must be kept above a certain threshold, or the face is unstable
and there is risk of tunnel collapse.

To prevent this it is necessary to support soft and non-cohesive soils from the time they
are excavated to the moment the final support is installed. Where groundwater is present
it is also necessary to prevent a flow towards the tunnel face, as this flow may have an
eroding effect on the tunnel face. In a tunnel boring machine (TBM) drive, the radial
support and water tightness is initially ensured by the shield and then by the tunnel
lining. At the face, a mixture of the excavated soil mass and varying additives (or
pressurized slurry in case of a slurry TBM) are used to provide necessary face pressure.

Though many metro projects have been completed, due to the complicated geological
conditions tunnel face collapse still takes place occasionally in recent tunnelling project
[3]. This calls for a better understanding of the problem and the uncertainties involved
in it.
Although a number of studies have been dedicated to tunnel face stability, only the
collapse failure mode of the ultimate limit stated is studied. The deterministic model is
either based on the upper bound or lower bound methods. The inherent uncertainty
(spatial and in determination of properties) are not considered in the deterministic
models.

1
1.2 Objective and Scope of the Study

A number of studies have been dedicated to tunnel face stability[1] and recent
developments (like “multi-block mechanism”, “method of slices” etc.) have updated the
widely used methods of analysis. Most analytical results are based on the limit
equilibrium method and the limit analysis method. This study is focused to discuss the
basis and use of different methods for different practical conditions from a design office
point of view.
Numerical methods were also frequently used to investigate stability of the tunnel face.
However, because of the level of attention required to select proper constitutive model
and parameters, numerical analyses could be time consuming. Moreover, the validity of
numerical analyses should also be checked, either by in situ measurements or by
laboratory model tests. Hence numerical analysis are generally difficult to be adopted
for routine design [4] and hence not discussed extensively in this report.
More importantly, the implementation of probabilistic analysis for computing
probability of face collapse using the analytical method is explored. This will allow to
incorporate uncertainty and variability of geotechnical parameters considered for the
analysis by performing probability analysis of failure. Similar techniques are already a
standard procedure for geotechnical problems like slope stability etc. Simplified model
for implementing probabilistic analysis with various face stability analytical models are
explored and presented in this study for quick implementation in design routines.

1.3 Organization of the report

Chapter 2 deals with different tunnel stability models reported in the literature. Within
the chapter, the stability models are classified based on the soil-type, briefly illustrated
and compared. Developments and recent trends in the stability models are explained
and compared with each other. Chapter 3 deals with the Probabilistic estimate of face
support pressure. Methods to compute probability of failure and statistical reliability
quantification for the design are discussed in this chapter. Chapter 4 presents the
observations and summarizes the work presented in this report.

2
2 Tunnel Face Stability

One of the main objectives of the tunnel boring process is to adequately support the soil
and to minimize deformations during and after construction. The required face pressure
is maintained to ensure safe working conditions. It should not be as low as to allow
uncontrolled collapse of the soil into the working chamber, nor as high as to lead to large
deformations of the soil or to a blow-out or in some cases, change in pore water pressure.
The actual support pressure would mainly depends on encountered soil and groundwater
conditions, the excavation method and the size and overburden of the tunnel. To find
the minimal and maximal allowable support pressures, a number of models has been
proposed in literature over the years to describe different possible failure mechanisms of
the tunnel face and to calculate the properties of the support medium necessary to
prevent collapse. In this chapter around 30 of such stability models are discussed and
briefly illustrated. Section 2.3 presents a comparison of limit support pressure values
obtained using some of the discussed stability models.

For the estimation of support pressure, there are mainly two types of methods which are
commonly used to assess the proper support pressure:
• Limit Equilibrium Analysis
• Limit State Concept – Stress method based on plasticity

The limit equilibrium solutions including the Silo theory [5] are improved by other
researchers, for example,. Anagnostou and Kovari [6], [7] for homogenous soil and later
for heterogeneous soils by Broere [8], Anagnostou [9], [10] etc.
Using the limit state concept, Davis et al. [11] investigated the support pressure for
tunnelling in cohesive soil by a 2D plane analysis. Afterwards, Muelhaus [12] found a
lower bound solution for the tunnel stability in frictional soil characterized by the Mohr–
Coulomb criterion. Leca and Dormieux [13] improved the former lower bound solution
and carried out an upper bound analysis for the tunnel face stability in low- cohesive

3
soil in 3D. It is further developed recently to multi-block failure mechanism by Mollon
[14]. Tang et al [3] extended the limit state concept for layered soils.
It is evident from above summary that the oldest stability model dates back to 1961 and
many recent models are being developed to address the heterogeneity in the soil and
actual difficult geological conditions. Following sections discusses the stability models,
classified based material model and results are compared.

2.1 Face Stability Models

When discussing previous research on tunnel heading stability, one has to distinguish
between drained and undrained conditions. For undrained conditions, as dominant in
clays, practical design curves have been derived on the basis of model tests [15] and these
curves have been largely confirmed by theoretical studies [11]. The question whether a
drained or undrained stability analysis should be carried out can be answered by
considering the type of ground and the advance rate of the tunnel face. According to a
parametric study by Anagnostou and Kovári [2], drained conditions tend to apply when
the ground permeability is higher than 10-7 to 10-6 m/s and the net excavation advance
rate is 0.1 to 1.0 m/hr or less. In a predominately sandy soil, therefore, drained stability
conditions should be considered. In a clayey, low-permeability soil the undrained analysis
is valid during excavation, but the drained analysis applies in case of a standstill. Hence,
even for excavations in clay it is important to investigate drained soil conditions.
It is also observed that from the laboratory tests that the geometry of the failure
mechanism in sand and in clay is notably different. Tests in sand show a chimney-like
failure mechanism, described by the wedge stability models. Centrifuge tests on clay
show a much larger and widening zone influenced by the instability Figure 1.
Hence, this chapter briefly presents the stability models classified based on the soil model
for which they are applicable. List of all stability models studied in this report is enclosed
in Appendix A. This classification can be useful to select appropriate model for a given
design consideration.

4
Figure 1. Failure mechanism observed in sand and clay [16]

2.1.1 Models for Undrained Cohesive Soil (Tresca)

This section discusses the stability models available in literature specifically for
undrained cohesive soft soils. Summary of the stability models discussed in this section
are presented in table below.
Table 1. Summary of stability models for cohesive soil
Name of the Model Year Soil Type Analysis Method Reference

Broms and Bennemark 1967 Homogeneous Empirical [17]

Davis et al 1980 Homogeneous Stress Method [11]

Kimura & Mair 1981 Homogeneous Empirical [18]

Global Equilibrium /
Ellstein 1986 Homogeneous [19]
Limit Equilibrium

Mori & Bezuijen 1991, 96 Homogeneous Empirical [20], [21]

Stress Method (using


Augarde et al 2003 Homogeneous [22]
FEM Simulations)

Klar et al 2007 Homogeneous Stress Method [23]

5
Stress Methods are analytical methods, which are based on the construction of statically,
and kinematically admissible solutions to take advantage of what are known the lower
and upper bound theorems of plasticity.

B rom s and B ennem ark (1967)

Broms and Bennermark [17] was one of the first stability models for cohesive soils. This
method provides a relation for the stability analysis of an unsupported opening in
cohesive undrained material - Tresca criteria,[24]. With this relationship they define the
stability ratio N to be equal to the difference between total overburden stress and
support pressure divided by the undrained shear strength

𝑁𝑁 = (𝑞𝑞𝑠𝑠 − 𝜎𝜎𝑇𝑇 )/𝑐𝑐𝑢𝑢 + (𝐶𝐶 + 𝑅𝑅). 𝛾𝛾/𝑐𝑐𝑢𝑢

where,
γ is soil density, cu is undrained cohesion, C is overburden, R is tunnel radius, qs is surface
load. From observations of collapses in both building pits and tunnel constructions, and
from laboratory extrusion tests, they find that an opening in such conditions will become
unstable N ≥ 6.

𝜎𝜎 = 𝛾𝛾. (𝐶𝐶 + 𝑅𝑅) + 𝑞𝑞𝑠𝑠 − 𝑁𝑁. 𝑐𝑐𝑢𝑢

Figure 2. Broms and Bennemark face stability model [17]

6
D avis et al (1980)

In 1980 Davis et al. [11] investigate the stability of an idealised partially unlined tunnel
heading in a Tresca material and introduce the distance P between the face and the
point where a stiff support is provided. They use the vertical opening as presented by
Broms & Bennermark as one of three limit cases for their stability analysis. They were
perhaps the first to propose that a statically admissible solution for a spherical opening,
could be used to analyse the stability of the front of a lined tunnel [25]. They further
derive upper and lower bound plasticity solutions for the stability of a plane strain
unlined cavity and a plain strain ‘long wall mining’ problem, obtained by taking P =
Infinity.

Figure 3. Davis' face stability model (a) [11]

Figure 4. Davis' face stability model (b) [11]

7
Of particular relevance to the face stability of a tunnel excavated using a shield machine
is the case P = 0. Davis et al. derive two lower bound solutions, using a cylindrical and
a spherical stress field.
𝐶𝐶
𝑁𝑁 = 2 + 2 ln � + 1� [𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐]
𝑅𝑅

𝐶𝐶
𝑁𝑁 = 4 . ln � + 1� [𝑠𝑠𝑠𝑠ℎ𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒]
𝑅𝑅

K im ura and M air (1981)

In 1981, Kimura and Mair [18] conducted a number of Cambridge centrifuge and
obtained a stability ratio design line for tunnels in undrained conditions. When plotted,
the results form neat hyperbolic patterns which are easily analysed.

Figure 5. Stability ratio design line proposed by Kumara & Mair [18]

Ellstein (1986)

Ellstein [19] gave an analytical expression of N for homogeneous cohesive soils based on
a limit equilibrium analytical approach. His results are in good agreement with those by
Kimura and Mair [18].

8
Figure 6. Ellstein's face stability model [19]

For a = 0, he proposed a stability number Nc as:

4 + √2
𝑁𝑁𝑐𝑐 = 𝐷𝐷
𝐾𝐾° + 1.5𝐻𝐻

For the tunnels excavated using fluid pressure against heading, Nc is modified as:

4 + √2
𝑁𝑁𝑐𝑐 = 𝐷𝐷 𝑝𝑝
𝐾𝐾° + 1.5𝐻𝐻 − 𝛾𝛾𝛾𝛾

Where p being the applied fluid pressure, H being the overburden above crown, D is
outer diameter of the tunnel, h is H+D/2 and K0 is coefficient of earth pressure at rest.

M ori & B ezuijen (1991 & 1996)

Mori [20] and Bezuijen [21] have come up with an expression for fracturing model to
estimate blow-out pressure.
Mori [20] defines the pressure at which (vertical) fracturing must occur for a normally
consolidated soil as
𝑠𝑠𝑓𝑓 = 𝐾𝐾0 𝜎𝜎𝜗𝜗′ + 𝑝𝑝 + 𝑞𝑞𝑢𝑢

with qu the unconfined compressive strength. In general fracturing will occur somewhat
earlier, resulting in an unsafe estimate of the maximal support pressure. For highly over

9
consolidated soils with K0 ≥ 1, the horizon
tal effective stress may be greater than the
vertical, as a result of which horizontal fracturing will occur at a stress significantly
lower than determined above.

Bezuijen [21] states a more general formula,

𝑠𝑠𝑓𝑓 = 𝜎𝜎3 + η𝑓𝑓 𝑐𝑐𝑢𝑢

which defaults to Mori's expressed for ηf = 2 and normally consolidated soil. Bezuijen
holds that when taking ηf = 1, this expression will yield safe estimates of the maximum
allowable support pressure.

A ugarde et al (2003)

Augrade [22] presented the stability of an idealised heading in undrained soil conditions.
They presented load parameters, for a wide range of heading configurations and ground
conditions using Finite element limit analysis methods, based on classical plasticity
theory. The approach presented by Augarde et al uses load parameter (σs - σT) / cu0
instead of N (stability ratio or overload factor) and also accounted for cu varying with
the depth. The results of the FE bound analysis are presented as dimensionless stability
charts and tables for easy implementation in design routine. The load parameter table
is included here for quick reference.

Figure 7. Layout of plane strain heading problem [22]

10
Table 2. Load parameter table based on [22]

K lar et al (2007)

Klar et al [23] suggest an upper bound solution for cohesive soils. They adopted elasticity
theory to estimate an admissible velocity field over which differential equation of
continuity are integrated. The velocity field for the upper bound solution is assumed to
be proportional to the displacement field.

11
Figure 8. Failure mechanism proposed by [23]

The stability number follows the definition of Davis et al. [11]

𝑁𝑁 = [𝜎𝜎𝑠𝑠 − 𝜎𝜎𝑡𝑡 + 𝛾𝛾(𝐶𝐶 + 𝐷𝐷/2)]/𝑠𝑠𝑢𝑢

where C is the tunnel cover depth and D is the tunnel diameter. If an upper bound value
is evaluated for σs - σT, then

̇
∫ 𝑉𝑉2�∈|𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 + ∫ 𝑆𝑆3 𝑣𝑣𝑡𝑡 𝑑𝑑𝑑𝑑 − (𝛽𝛽/𝐷𝐷) ∫ 𝑉𝑉 𝑣𝑣𝑣𝑣 𝑑𝑑𝑑𝑑 𝐶𝐶 + 𝐷𝐷/2
𝑁𝑁 = + 𝛽𝛽
𝑉𝑉 . 𝐿𝐿 𝐷𝐷

where β= γ D / Cu and the variation with respect to C/D ratio is presented in figure
below.

12
Figure 9. Stability number for 2D case [23]

2.1.2 Models for Cohesionless Soil

Summary of the stability models discussed in this section are presented in Table 3.
Although models available for Mohr-coulomb material could be easily used for pure
cohesionless soil, this classification is made to capture the development of stability
models specifically for cohesionless soils.

Table 3. Summary of stability models for cohesionless Soil


Name of the Model Year Soil Type Analysis Method Reference

Atkinson and Potts 1977 Homogeneous Stress method [26]

Chambon and Corte 1994 Homogeneous Stress method [27]

Chen et al 2014 Heterogeneous Global equilibrium [28]

A tkinson and Potts (1977)

In 1977, Atkinson & Potts [26] derived the minimal support pressure for an unlined
cavity in a dry cohesionless material. They differentiate between two limit cases. The

13
first case is a tunnel in a weightless medium with a surface load, the second a tunnel in
a medium with γ>0 but without surface loading. For the second case, two lower limit
solutions are furnished. The solution, which is independent of the overburden, provides,
in general, the result associated with the greater safety:

𝑠𝑠𝑚𝑚𝑚𝑚𝑚𝑚 = �2𝑘𝑘𝑝𝑝 /�𝑘𝑘𝑝𝑝2 − 1�� × 𝛾𝛾 × 𝑅𝑅

Where,
𝑘𝑘𝑝𝑝 = (1 + sin 𝜑𝜑)/(1 − sin 𝜑𝜑)

As this represents a statically admissible stress field, it is a safe estimate for the minimal
support pressure. The kinematic upper bound solution on the other hand, which is the
inherently unsafe estimate, yields a lower value for the minimal support pressure. It
should be noted that both upper and lower bound solutions found by Atkinson & Potts
are independent of the relative overburden C/D.

Figure 10. Stability model of the method Atkinson and Potts [26]

Cham bon and Corte (1994)

Chambon & Corté [27] used a model tunnel sealed by a membrane and embedded in a
homogeneous sand layer. They gradually reduced the support pressure and found wedge
shaped failure planes loaded by soil silos. Their investigation focussed on the influence
of the overburden on the silo formation and shows that a soil silo will form even if only
a small overburden is present but that its influence will be limited by the proximity of

14
the soil surface. They further show that a soil silo can develop fully at an overburden
equal to the tunnel diameter or larger. Chambon and Corte also found that the cover-
to-diameter ratio C/D (i.e., relative depth; C is the cover depth and D is the diameter
of the tunnel) has significant effect on the failure mechanism. That means, when relative
depth C/D is low (e.g., C/D = 0.5), the failure zone in the limit state has extended to
the ground surface. When C/D is high (e.g., C/D = 1 or 2), the failure zone in the limit
state will be still in the interior of the ground.

Figure 11. Failure bulbs comparison with theoretical mechanism [27]

The test results have provided better understanding of the behaviour of tunnel face in
pure cohesion less soils and have suggested that the process leading to failure of face
could be (p is pressure applied, pc is the pressure at which the first movements of the
face could occur and pf is the pressure at which failure would occur:
For p> pc, no movement of face would occur
For pc > p > pf, small displacements of the face and surface settlements would occur
For p = pf, there would be a sudden but localized collapse
For p < pf, there would be a flow of soil in the tunnel

There is no time effect in any of the stages; at constant internal pressures, the tunnel
face is stabilized. In a dry sand, in all cases, including in the presence of a load on the

15
surface, a hydrostatic pressure is sufficient to provide stability to the face. However, the
face is not self-stabilizing and must be supported. Small uniform pressures, of the order
of about 10 kPa, are sufficient. The pressure at failure is little affected by changes of
geometry at the relative depths considered (C/D = 0.5, 1, 2, 4) and by the density of
the soil.

Chen et al (2014)

Chen et al. [28] proposed a 3D wedge prism model which is based on limit equilibrium
method to determine the limiting the support pressure at the tunnel face. This model
considers the influence of relative depth on the failure mechanism as observed by
Chambon and Corte [27]. The other aspect of the improvement is properly considering
soil arching effects (Broere, [8]) to calculate the vertical force applied on the wedge’s
upper surface more accurately, as it has significant influence on the limit support
pressure.

Figure 12. Improved 3D wedge prism model [28]

16
Figure 13. Mechanical Equilibrium of the wedge [28]

Chen et al. proposed expression for limiting support pressure:

4𝑆𝑆
𝑠𝑠𝑙𝑙𝑙𝑙𝑙𝑙 = 𝑚𝑚𝑚𝑚𝑚𝑚 � � = 𝑚𝑚𝑚𝑚𝑚𝑚{𝑓𝑓1 𝜎𝜎𝑎𝑎𝑎𝑎 + 𝑓𝑓2 𝛾𝛾𝛾𝛾}
𝜋𝜋𝐷𝐷2

Where,
tan(𝛽𝛽 − 𝜙𝜙)
𝑓𝑓1 = − 𝜆𝜆ℰ tan 𝜙𝜙
tan 𝛽𝛽
tan(𝛽𝛽 − 𝜙𝜙) 1
𝑓𝑓2 = − 𝜆𝜆ℰ tan 𝜙𝜙
2 tan 𝛽𝛽 3
4
𝜀𝜀 = [cot 𝛽𝛽 + tan(𝛽𝛽 − 𝜙𝜙)] cos 𝛽𝛽
𝜋𝜋
𝜋𝜋 𝜋𝜋
𝛽𝛽𝛽𝛽 � , �
4 2

Where, λ is the ratio of horizontal to vertical stress in the wedge given by:
𝜆𝜆 = 1 − sin Φ

𝛾𝛾𝛾𝛾 𝑈𝑈
𝑠𝑠 tan 𝜙𝜙𝜙𝜙
−𝐾𝐾 𝐴𝐴
𝜎𝜎𝑎𝑎𝑎𝑎 = �1 − 𝑒𝑒 �
𝑈𝑈𝐾𝐾 𝑠𝑠 tan 𝜙𝜙

17
with,
𝜋𝜋𝜋𝜋 𝐷𝐷
𝐴𝐴 = 𝐵𝐵 × 𝐿𝐿 = ×
4 tan 𝛽𝛽
𝜋𝜋𝜋𝜋 𝐷𝐷
𝑈𝑈 = 2 × (𝐵𝐵 + 𝐿𝐿) = 2 × � + �
4 tan 𝛽𝛽
2𝐷𝐷
𝐻𝐻 = 𝑚𝑚𝑚𝑚𝑚𝑚 �𝐶𝐶, �
tan 𝛽𝛽
2𝐷𝐷
𝑞𝑞𝑜𝑜 = 𝛾𝛾 × �𝐶𝐶 − 𝑚𝑚𝑚𝑚𝑚𝑚 �𝐶𝐶, ��
tan 𝛽𝛽

Figure 14. Distribution of the vertical stress σz along side slip surfaces [28]

𝑐𝑐𝑐𝑐𝑐𝑐 2 𝜃𝜃𝑜𝑜 + 𝐾𝐾𝑎𝑎 𝑠𝑠𝑠𝑠𝑠𝑠2 𝜃𝜃0


𝐾𝐾 𝑠𝑠 = 1
(1 − 𝐾𝐾𝑎𝑎 )𝑠𝑠𝑠𝑠𝑠𝑠2 𝜃𝜃𝑜𝑜 + 𝐾𝐾𝑎𝑎
3

where
𝜋𝜋 𝜙𝜙 𝜋𝜋 𝜙𝜙
𝜃𝜃𝑜𝑜 = + , 𝐾𝐾𝑎𝑎 = 𝑡𝑡𝑡𝑡𝑡𝑡2 � − �
4 2 4 2

This model has sufficient accuracy as the limit support pressure obtained via this model
agrees well with that obtained via Vermeer et al.’s method [29] which was proposed
based on a series of rigorous FEM calculations. In addition, the improved wedge-prism
model is also a relatively safe estimation of the limit support pressure for the practical
application.

18
0.25
C/D = 1
C/D = 0.5
0.2 C/D = 2
C/D = 3
C/D = 4
0.15
Slim/γ.D

0.1

0.05

0
25 30 35 40 45
Phi [degrees]
Figure 15. Design chart by Chen et al for cohesionless soils

Chen et al also presented a design chart (Figure 15) for easy implementation in design
office.

19
2.1.3 Models for C-Phi Soil (Mohr Coulomb)

Table 4. Summary of stability models for C-Phi Soil


Name of the Model Year Soil Type Analysis Method Reference

Murayama 1966 Homogeneous Global equilibrium [30]

Mühlhaus 1985 Homogeneous Stress Method [12]

Krause 1987 Homogeneous Global equilibrium [31]

Mohkam 1989 Homogeneous Global equilibrium [32]

Leca and Dormieux 1990 Homogeneous Stress method [13]

Jancsecz & Steiner 1994 Homogeneous Global equilibrium [33]

Anagnostou &Kovári 1996 Homogeneous Global equilibrium [2]

Belter and Katzenbach 1999 Heterogeneous Global equilibrium [30]

Broere 2001 Heterogeneous Global equilibrium [8]

Vermeer & Ruse 2001 Homogeneous Semi-empirical [34]

Caquot’s Model (Carranza-Torres) 2004 Homogeneous Stress method [25]

Kolymbas 2005 Homogeneous Stress method [35], [36]

Girmschied 2005 Homogeneous Global equilibrium [35]

Mollon et al 2010 Heterogeneous Stress method [14]

Anagnostou 2012 Heterogeneous Global equilibrium [9]

Perazzeli, Leone and Anagnostou 2014 Heterogeneous Global equilibrium [10]

M urayam a (1966)

Murayama (as cited in [30]) calculated the minimal support pressure using a two-
dimensional log-spiral shaped sliding plane in 1966. The soil weight acting on the
pressure wedge is calculated in accordance with the Terzaghi (1943) theory. The face
stability requires the equilibrium between the moment of the acting weight forces and
the resistant forces i.e, the force applied on the tunnel face (P) and shear strength along
the failure surface. The method contemplates the iterative search for the solid-load width

20
that determines the more unfavourable loading condition and, therefore, the maximum
stabilisation pressure, P.

1 𝜔𝜔1 2
𝑟𝑟𝑎𝑎2
𝑠𝑠𝑚𝑚𝑚𝑚𝑚𝑚 = (𝐺𝐺𝐼𝐼 )
𝐺𝐺 + 𝑞𝑞𝑤𝑤 𝜔𝜔1 �𝑙𝑙𝜔𝜔 + � − 𝑐𝑐 �𝑟𝑟𝑑𝑑 − �
2𝑅𝑅𝑙𝑙𝑝𝑝 2 2 tan 𝜑𝜑

Figure 16. Failure mechanism by Murayama [8]

M ühlhaus (1985)

Mühlhaus [12] was first to propose a complete statically admissible solution for tunnel
stability. The solution could be used to analyse the stability of the unsupported span in
the vicinity of the tunnel front. Since in urban TBM drive, unsupported heading is
almost zero, this model is not relevant for face pressure calculation and hence not
discussed further.

21
Figure 17. Statically admissible solution by Muehlhaus [12]
K rause (1987)

Krause [31] proposed minimal support pressures needed for a semi-circular and spherical
limit equilibrium mechanism using the shear stresses on the sliding planes. Of the three
mechanisms proposed, the quarter circle (Figure 18 b) will always yield the highest
minimal support pressure. However, this may not always be a realistic representation of
the actual failure body. In many cases the half-spherical body (Figure 18c) will be a
better representation.

Figure 18. Failure mechanics proposed by Krause [8]

The highest minimal support pressure are given as:

For Half sphere,


1 1 1
𝑠𝑠𝑚𝑚𝑚𝑚𝑚𝑚 = � 𝐷𝐷𝛾𝛾 ′ − 𝜋𝜋𝜋𝜋�
𝑡𝑡𝑡𝑡𝑡𝑡 𝜑𝜑 9 2

22
For Half circle,
1 1 1
𝑠𝑠𝑚𝑚𝑚𝑚𝑚𝑚 = 1 � 𝐷𝐷𝛾𝛾 ′ − 𝜋𝜋𝜋𝜋�
2
+ 𝑡𝑡𝑡𝑡𝑡𝑡 𝜑𝜑 6 2

For Quarter circle,


1 1 1
𝑠𝑠𝑚𝑚𝑚𝑚𝑚𝑚 = � 𝐷𝐷𝛾𝛾 ′ − 𝜋𝜋𝜋𝜋�
𝑡𝑡𝑡𝑡𝑡𝑡 𝜑𝜑 3 2

M ohkam (1989)

Mohkam [32] described a limit equilibrium model using a roughly log-spiral shaped wedge
which has to be obtained from a variational analysis over the unknown position angle
and total stress of the failure plane. The model also includes the effect of the reduced
effectiveness of the slurry pressure due to infiltration using a non-linear relation between
infiltration distance and effective slurry pressure.
Taking into account the support-free length before the installation of a stiff support, two
failure mechanisms are assumed: one involves the face excavation (Figure 19 a & b) and
the other involves the tunnel wall (Figure 19c), along the failure surface, respectively,
logarithmic spiral and cylindrical. The load acting on the wedge is based on Terzaghi’s
arch effect. The large number of unknowns present in the model leads to a highly
iterative solution procedure and a rather unwieldy model and hence not discussed further
in this report.

Figure 19. Failure mechanics proposed by Mohkam [1]

23
Leca and D orm ieux (1990)

Leca and Dormieux [13] propose a series of conical bodies in 1990 (Figure 22). Combined
with different stress states, similar to those proposed by Davis et al., they derive lower
and upper bound limits for both the minimal and maximal support pressure of a limed
tunnel in a dry Mohr-Coulomb material. Three failure mechanisms which involve the
movement of solid conical blocks with circular cross sections, are shown in Figure 21.
MI and MII failure mechanisms are single-cone and dual-cone systems, respectively;
where the cones move into the excavation. An MIII failure mechanism is a single-cone,
passive mechanism, where the cone moves outward to the surface.

Figure 20. Conical blocks and kinematic conditions used in MI, MII and MIII [13]

Figure 21. 3D visualization of failure mechanism by Leca and Dormieux [37]

24
Figure 22. Failure mechanics proposed by Leca and Dormieux [13]

Expression for support pressure in MII collapse failure mechanism, which is most
common failure pattern in shield tunnelling [27], is presented below:

𝜎𝜎𝑠𝑠 𝛾𝛾𝛾𝛾 𝜎𝜎𝑇𝑇


𝑁𝑁𝑠𝑠 ��𝐾𝐾𝑝𝑝 − 1� + 1� + 𝑁𝑁𝛾𝛾 �𝐾𝐾𝑝𝑝 − 1� ≤ �𝐾𝐾𝑝𝑝 − 1� + 1
𝜎𝜎𝑐𝑐 𝜎𝜎𝑐𝑐 𝜎𝜎𝑐𝑐
Where,

1 sin(𝛽𝛽 − 𝜙𝜙 ′ )𝑅𝑅𝐸𝐸2
𝑁𝑁𝑠𝑠 =
cos 𝛼𝛼 𝑐𝑐𝑐𝑐𝑐𝑐 2 𝜙𝜙 ′ sin(𝛽𝛽 + 𝜙𝜙 ′ )𝑅𝑅𝐴𝐴

25
1 𝑐𝑐𝑐𝑐𝑐𝑐𝜙𝜙 ′ cos(𝛽𝛽 + 𝜙𝜙 ′ ) 𝑅𝑅𝑐𝑐3 1 sin(𝛽𝛽 − 𝜙𝜙 ′ ) 𝑅𝑅𝐸𝐸3
𝑁𝑁𝛾𝛾 = [𝑡𝑡𝑡𝑡𝑡𝑡𝑅𝑅𝐵𝐵 + − ×
3 2𝑠𝑠𝑠𝑠𝑠𝑠𝜙𝜙 ′ cos(𝛽𝛽 + 𝜙𝜙 ′ ) 𝑅𝑅𝐴𝐴 2𝑠𝑠𝑠𝑠𝑠𝑠𝜙𝜙 ′ cos 𝛼𝛼𝑐𝑐𝑐𝑐𝑐𝑐 2 𝜙𝜙 ′ sin(𝛽𝛽 + 𝜙𝜙 ′ ) 𝑅𝑅𝐴𝐴

Figure 23. Geometry of Mechanism MII [13]

�[cos(𝛼𝛼 − 𝜙𝜙 ′ ) cos(𝛼𝛼 + 𝜙𝜙 ′ )]
𝑅𝑅𝐴𝐴 =
cos 𝜙𝜙 ′

[cos(𝛼𝛼 − 𝜙𝜙 ′ ) cos(𝛼𝛼 + 𝜙𝜙 ′ )]
𝑅𝑅𝐵𝐵 =
sin(2𝜙𝜙 ′ )

1⁄2
cos(𝛼𝛼 + 𝜙𝜙 ′ ) sin(β − ϕ′ )
𝑅𝑅𝐶𝐶 = � �
cos 𝜙𝜙 sin(𝛽𝛽 + 𝜙𝜙 ′ )
sin 𝛽𝛽
𝑅𝑅𝐷𝐷 =
sin 𝜙𝜙 ′ sin(𝛽𝛽 + 𝜙𝜙 ′ )

𝑐𝑐𝑐𝑐𝑐𝑐 2 𝜙𝜙 ′ 2𝐶𝐶
𝑅𝑅𝐸𝐸 = 𝑅𝑅𝑐𝑐 − sin 𝜙𝜙 ′
cos(𝛼𝛼 + 𝜙𝜙 ′ ) 𝐷𝐷

26
Above relation provides the best upper bound association with MII when α chosen such
that Ns , and Nγ are at minimum. The above results only apply when the ground surface
is reached by the failure mechanism. that is to say when:

𝐶𝐶 cos(𝛼𝛼 + 𝜙𝜙 ′ ) sin(𝛽𝛽 − 𝜙𝜙 ′ )

𝐷𝐷 2 sin 𝜙𝜙 ′ sin(𝛽𝛽 + 𝜙𝜙 ′ )

For deeper tunnels, RE is set to zero in above relations.

Recently, this method was updated by Lee et al. [38]to account for seepage forces acting
on the tunnel face under steady-state flow conditions. However, general solutions which
combine the depth-dependence of the effective cohesion of NC clays and the influence of
seepage have not been reported.
Expression for support pressure in MIII blow-out failure mechanism is presented below:

Figure 24. Geometry of Mechanism MIII [13]

27
𝑅𝑅𝐴𝐴 = cos 𝛼𝛼 �[cos(𝛼𝛼 + 𝜙𝜙 ′ ) cos(𝛼𝛼 − 𝜙𝜙 ′ )]

𝑅𝑅𝐵𝐵 = sin 𝛼𝛼 �[sin(𝛼𝛼 + 𝜙𝜙 ′ ) sin(𝛼𝛼 − 𝜙𝜙 ′ )]

sin 2𝛼𝛼 + (2𝐶𝐶 ⁄𝐷𝐷 + 1) sin 2𝜙𝜙 ′


𝑅𝑅𝑐𝑐 =
cos 2𝜙𝜙 ′ − cos 2𝛼𝛼

𝜎𝜎𝑠𝑠 𝛾𝛾𝛾𝛾 𝜎𝜎𝑇𝑇


𝑁𝑁𝑠𝑠 ��𝐾𝐾𝑝𝑝 − 1� + 1� + 𝑁𝑁𝛾𝛾 �𝐾𝐾𝑝𝑝 − 𝑡𝑡� ≥ �𝐾𝐾𝑝𝑝 − 1� + 1
𝜎𝜎𝑠𝑠 𝜎𝜎𝑐𝑐 𝜎𝜎𝑐𝑐

𝑅𝑅𝐵𝐵 𝑅𝑅𝑐𝑐2
𝑁𝑁𝑠𝑠 =
𝑅𝑅𝐴𝐴
sin 𝛼𝛼 𝑅𝑅𝐵𝐵 𝑅𝑅𝐶𝐶 3 𝑅𝑅𝐴𝐴 3
𝑁𝑁𝛾𝛾 = �� � −� � �
3𝑅𝑅𝐴𝐴 sin 2𝜙𝜙1 sin 𝛼𝛼 cos 𝛼𝛼

The best upper bound associated with MIII is found by choosing a such that
Ns , and Nγ, are at minimum. Since failure always reaches the ground surface, above
expression is valid for all values of C/D.

Jancsecz & Steiner (1994)

Jancsecz & Steiner [33] proposed a soil wedge model similar to Horn [5], Method. The
three-dimensional failure scheme shown in Figure 25 consists of a soil wedge (lower part)
and a soil silo (upper part). The vertical pressure resulting from the silo and acting on
the soil wedge is calculated according to Terzaghi’s solution.

28
Figure 25. Stability model by Jancsecz & Steiner [33]

The three dimensional earth pressure coefficient is given as:

𝑘𝑘𝑎𝑎3 = (sin 𝛽𝛽 . cos. −𝑐𝑐𝑐𝑐𝑐𝑐 2 𝛽𝛽. tan 𝜙𝜙 − 𝐾𝐾. 𝛼𝛼. cos 𝛽𝛽. tan 𝜙𝜙⁄1.5)⁄(sin 𝛽𝛽. cos 𝛽𝛽 + 𝑠𝑠𝑠𝑠𝑠𝑠2 𝛽𝛽. tan 𝜙𝜙)

where,
Κ ≈ [1 − sin ϕ + tan2 (45 + ϕ⁄2)]⁄2 ; α = (1 + 3 . t⁄D)⁄(1 + 2 . t⁄D).

and limiting support pressure is given as:


𝑠𝑠𝑚𝑚𝑚𝑚𝑚𝑚 = 𝐾𝐾𝑎𝑎3𝐷𝐷 𝜎𝜎𝜗𝜗′ + 𝑝𝑝

A nagnostou & K ovári (1996)

Anagnostou & Kovári [2] method, is based on the silo theory [33] and to the three-
dimensional model of sliding mechanism proposed by Horn [5]. The analysis is performed
in drained condition, and a difference between the stabilizing water pressure and effective
pressure in the plenum of a EPBs is presented. If there is a difference between the water
pressure in the plenum and that in the ground, destabilizing seepage forces occur and a
higher effective pressure is required at the face.

29
Figure 26. Seepage force f and effective support pressure s’ [2]

However, accepting this flow, the total stabilizing pressure is lower than the pressure
required in the case of an imposed hydro-geological balance. The effective stabilizing
pressure is:

𝜎𝜎 ′ = 𝐹𝐹𝑜𝑜 . 𝛾𝛾 ′ . 𝐷𝐷 − 𝐹𝐹1 . 𝑐𝑐 ′ + 𝐹𝐹2 . 𝛾𝛾 ′ . Δℎ − 𝐹𝐹3 . 𝑐𝑐 ′ . Δℎ/𝐷𝐷

where F0, F1, F2, F3are non-dimensional factors derived from nomograms, which are
function of H/D and friction angle.

Figure 27. Stability model of Anagnostou & Kovari based on Horn [2]

30
Figure 28. Nomograms for non-dimensional factors F0, F1 [2]

Figure 29. Nomograms for non-dimensional factors F2, F3 [2]

B elter and K atzenbach (1999)

Belter and Katzenbach introduced for the first time, a two-dimensional wedge model
that can include layered soils above the TBM, but not in front of the TBM [30]. It
remains unclear from the brief description, however, whether this model includes arching
effects of the soil above the tunnel or to what extent these heterogeneities influence the
required support pressure. The model sketched by Belter and Katzenbach (shown in
Figure 30) could also be obtained by including layered soil above the face in the wedge

31
models described by Jancsecz & Steiner or Anagnostou & Kovári and hence not discussed
in detail here.

Figure 30. Stability model of Belter and Katzenbach [8]


B roere (2001)

Broere [8] pointed out some important shortcomings of the existing edge stability models
and developed a model to address the following aspects:

• Layer boundaries / heterogeneity of the ground at the face.


• Accounting soil-arching effect instead of presuming a linear stress distribution in
the evaluation of the vertical load.
• Reduction in effective support pressure due to permeability of the soil skeleton

Figure 31. Multilayer wedge model by Broere [8]

32
Figure 32. Multilayer wedge model by Broere [8]

For the layer “i” with top z = t (1) the following formulation is proposed for a stratified
soil, in the range t (i) < z < t (i+1):
′(𝑖𝑖) 𝑎𝑎𝛾𝛾 ′(𝑖𝑖) − 𝑐𝑐 ′(𝑖𝑖) ′(𝑖𝑖)
−𝐾𝐾(𝑖𝑖) 𝑡𝑡𝑡𝑡𝑡𝑡𝑓𝑓 𝑎𝑎
𝑍𝑍
𝜎𝜎𝜈𝜈,𝑎𝑎 = (𝑖𝑖) �1 − 𝑒𝑒 �
𝐾𝐾 tan 𝜙𝜙 ′(𝑖𝑖)

It is found that for the simplified case of a single slide wedge in homogeneous soil, the
resultant formulation corresponds to that of Jancsecz [33].

V erm eer and R use (2001)

Vermeer and Ruse [34] conducted numerical analysis using FE-code PLAXIS and
deduced the following approximation for the limit support pressure for the case ϕ> 20°
𝑐𝑐 1
𝑝𝑝 ≈ − + 2𝛾𝛾𝛾𝛾 � − 0.05�
tan 𝜑𝜑 9 tan 𝜑𝜑

The underlying results are obtained with an elastic-ideal plastic constitutive law
assuming Mohr-Coulomb yield surface and associated plasticity. The consequences of
friction dependent arching are considerable. The stress arch carries the ground cover
independent of the magnitude of its thickness. Authors point that once the friction angle
is larger than about twenty degrees, stability is completely independent of the ground
cover. Compared to the soil weight stability number Nγ and the cohesion stability
number Nc, Nq is very low and it is disregarded by the authors.

33
For the case ϕ= 0, instead, they obtain a linear relation between p and h. For this
relation no analytic expression is given.

Caquot-K erisel m odified by Carranza-Torres (2004)

Caquot’s model considers the equilibrium condition for material undergoing failure above
the crown of a shallow circular (cylindrical or spherical) cavity. The material has a unit
weight γand a shear strength defined by Mohr-Coulomb parameters c (cohesion) and ϕ
(friction angle). A support pressure ps can be applied inside the tunnel, while a surcharge
qs acts on the ground surface. The Caquot generalised solution for dry conditions, can
be represented by the following equation developed by Carranza-Torres [25]

Figure 33. Extended Caquot's model [25]

𝑝𝑝𝑠𝑠 𝑞𝑞𝑠𝑠 𝑐𝑐 �𝑁𝑁𝜙𝜙 ℎ −𝑘𝑘�𝑁𝑁𝜙𝜙 −1� 1 ℎ −𝑘𝑘�𝑁𝑁𝜙𝜙−1� ℎ −1 𝑐𝑐 �𝑁𝑁𝜙𝜙


=� +2 �� � − �� � −� � �−2
𝛾𝛾ℎ 𝛾𝛾ℎ 𝛾𝛾ℎ 𝑁𝑁𝜙𝜙 − 1 𝑎𝑎 𝑘𝑘�𝑁𝑁𝜙𝜙 − 1� − 1 𝑎𝑎 𝑎𝑎 𝛾𝛾ℎ 𝑁𝑁𝜙𝜙 − 1

Where,
1 + sin 𝜙𝜙 𝜋𝜋 𝜙𝜙
𝑁𝑁𝜙𝜙 = = 𝑡𝑡𝑡𝑡𝑡𝑡2 � + �
1 − sin 𝜙𝜙 4 2

34
K olym bas (2005)

Kolymbas [35], [36] considered the distribution of the vertical stress between the ground
surface and the crown of a spherical cavity and approximated this distribution by a
quadratic parabola. Material strength is assumed to be fully mobilised at the crown.
Support pressure is given by the expression:
𝑐𝑐 cos 𝜑𝜑
𝛾𝛾 − 𝑟𝑟 1−sin 𝜑𝜑
𝑓𝑓
𝑝𝑝𝑐𝑐 = ℎ ℎ 2 sin 𝜑𝜑
1 + 𝑟𝑟 1−sin 𝜑𝜑
𝑓𝑓

Figure 34. Excavation face as hemisphere [36]

G irm scheid (2005)

Girmschied presented in Kirsch et al [35] presents another calculation procedure for


determination of face support pressure given by:

𝑐𝑐𝐷𝐷 2 𝐷𝐷 2 𝐷𝐷𝛾𝛾𝐵𝐵
+ 𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 �𝐾𝐾𝐾𝐾𝐾𝐾𝐾𝐾𝐾𝐾 �𝜎𝜎𝑣𝑣 (𝑡𝑡) + � + 𝑐𝑐�
𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 3
𝑆𝑆(𝜃𝜃) = (𝐴𝐴𝑉𝑉 + 𝐺𝐺) tan(𝜃𝜃 − 𝛿𝛿) + 𝑊𝑊 −
𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 (𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 + 1)

Where the maximum supporting pressure is to be selected by varying the values of θ.


The variables included in the function are:
Av is vertical force acting on the sliding wedge
G is the dead weight on the sliding wedge
θ is the angle of sliding surface

35
δ is effective wall friction angle of the sliding surface
D is the diameter of the excavation area
K is the earth pressure coefficient
σv(t) is vertical stress at depth

M ollon et al (2010)

Mollon et al [14] improved the existing limit analysis mechanisms by taking into account
the entire circular tunnel face and not only the inscribed ellipse. This was made using
spatial discretization technique and hence it is possible to generate multi-block
mechanism and not restricted to standard geometric shapes such as cones or cylinders.
Mollon et al [14] also presented design charts for easy implementation in design routines.
The critical values of Nγ, Nc and Ns presented in Figure 35 are using the superposition
method. It was shown that the error induced by the superposition principle is quite small
and is always conservative.

Figure 35. Multi-block failure mechanism [14]

Based on the non-dimensional coefficients, the face pressure is represented by:

𝜎𝜎𝑐𝑐 = 𝛾𝛾𝛾𝛾𝑁𝑁𝛾𝛾 − 𝑐𝑐𝑁𝑁𝑐𝑐 + 𝜎𝜎𝑠𝑠 𝑁𝑁𝑠𝑠

This method estimates the three dimensional surface by defining the contours at several
different vertical places parallel to the tunnel face. The failure mechanism respects the

36
normality condition required by limit analysis since the three dimensional failure surface
generated in such a manner that the velocity vector makes an angle ϕ with the velocity
discontinuity surfaces anywhere along these surfaces. Another advantage of this method
being, it can be applied to even tunnels of other cross sections.

Figure 36. Non-dimensional coefficients for face pressure estimation [14]

The proposed failure mechanism always outcrops in the case of purely cohesive soil. It
means that in this case the parameter C/D is of major importance. This is not the case
for c-phi soil which high to moderate friction angle (40 to 20 degree) since the critical
tunnel pressure is independent of the tunnel cover in these cases. Numerical results have
shown that multi-block mechanism composed of three blocks is a good compromise
between computation time and results accuracy.

37
0.4
Phi = 15
0.35 Phi = 20
Phi = 25
0.3
Phi = 30
0.25 Phi = 35
σc/γD

Phi = 40
0.2

0.15

0.1

0.05

0
0 0.02 0.04 0.06 0.08 0.1
c/γD
Figure 37. Design chart for critical collapse pressure for C-Phi soil [14]
A nagnostou (2012)

Anagnostou [9] revisits their computational model (Anagnostou and Kovári [7]) and an
alternative model is presented which is based on the method of slices. This method does
not need an a priori assumption as to the distribution of horizontal stress. This method
could accounts for horizontal arching around the tunnel face as the shear stresses τs
developing at the two vertical slip surfaces (as shown in Figure 38) contribute to the
stability of the wedge.

Figure 38. Cross section and horizontal section of a tunnel [9]

The method of slices is adopted form the stability assessment of slurry walls. In analogy
to the silo theory, the method of slices assumes proportionality between the horizontal
stress and vertical stress

38
𝜎𝜎𝑦𝑦′ = 𝜆𝜆𝜎𝜎𝑧𝑧′

where the coefficient of lateral stress λis assumed to be constant. In order to calculate
the distribution of the vertical stresses inside the wedge, the equilibrium of an
infinitesimally thin slice is considered (Figure 39). This makes it possible to analyse cases
with non-uniform face support, heterogeneous ground consisting of horizontal layers or
non-uniform distribution of the seepage forces along the height of the face.

Figure 39. Force acting on infinitesimal slice [10]

The method of slices also makes it possible to estimate on a more consistent basis
(similarly to silo theory) the vertical stresses within the wedge. The method of slices
leads to support pressures which are much closer to the numerical predictions of Vermeer
et al. [39]. These results indicate that the reason for the differences from the numerical
results is the simplified way of considering horizontal arching in the model of Anagnostou
and Kovári [6], [7]. The author also presented a simplified formula for estimation of
support pressure when λ = 1.0 (Terzaghi’s initial assumptions)

𝑠𝑠 = 0.05(cot 𝜙𝜙)1.75 𝛾𝛾𝛾𝛾 − cot 𝜙𝜙𝜙𝜙

Perazzeli, Leone and A nagnostou (2014)

In this model, Perazzeli et al [10], [40] combines the model for face stability under drained
condition by considering wedge and prism mechanism [6] with the method of method of
slices presented above [9]. They considered limit equilibrium model and the mechanism
fails if the load exerted by the prism upon the wedge exceeds the force which can be

39
sustained by the wedge at its upper boundary. At the limit equilibrium the prism load
is equal to the bearing capacity of the wedge. The prism load is calculated based on the
silo theory, while the bearing capacity of the wedge is calculated by considering the
equilibrium of the infinitesimal slice (as in [8]).

Figure 40. Stability model used for analysis [10]

The necessary support pressure is given by:

𝑠𝑠 ′ = 𝑚𝑚𝑚𝑚𝑚𝑚{(𝑠𝑠1′ 𝑜𝑜𝑜𝑜 𝑠𝑠2′ ), 𝑠𝑠3′ }

Where, s1' is the minimum support pressure required to ensure that tensile effective
stress does not exceed c'/tanϕ' (tensile strength exhibited by weak rocks) at any point
in the sliding surface and is given by:

𝑠𝑠1′ = 𝑃𝑃1 𝛾𝛾𝑤𝑤 Δℎ + 𝑃𝑃2 𝛾𝛾 ′ 𝐻𝐻 − 𝑃𝑃3 𝑐𝑐 ′

where,
1 − 𝑒𝑒 −𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏 Λ𝐶𝐶Δℎ (1)
𝑃𝑃1 = +
Λ𝐶𝐶𝑠𝑠 (1)
�1 + (𝑃𝑃 +cot 𝜔𝜔)� 𝑃𝑃𝑠𝑠 + cot 𝜔𝜔 + Λ𝐶𝐶𝑠𝑠 (1)
𝑠𝑠

Λ𝐶𝐶𝛾𝛾 (1)𝐵𝐵
𝑃𝑃2 =
(𝑃𝑃𝑠𝑠 + cot 𝜔𝜔 + Λ𝐶𝐶𝑠𝑠 (1))𝐻𝐻

40
𝑀𝑀𝑐𝑐 𝐻𝐻
𝐵𝐵
+ 𝑃𝑃𝑐𝑐 + Λ𝐶𝐶𝑐𝑐 (1) 𝑃𝑃𝑠𝑠 + cot 𝜔𝜔
𝑃𝑃3 = +
𝑃𝑃𝑠𝑠 + cot 𝜔𝜔 + Λ𝐶𝐶𝑠𝑠 (1) (𝑃𝑃𝑠𝑠 sin 𝜔𝜔 + cos 𝜔𝜔)(𝑃𝑃𝑠𝑠 + cot 𝜔𝜔 + Λ𝐶𝐶𝑠𝑠 (1)) tan 𝜙𝜙

s2' is the minimum support pressure required for soft soil without any tensile strength
and hence effective normal stress σn' must be higher than 0.
𝑠𝑠2′ = 𝑃𝑃1 𝛾𝛾𝑤𝑤 Δℎ + 𝑃𝑃2 𝛾𝛾 ′ 𝐻𝐻 − 𝑃𝑃4 𝑐𝑐 ′
where,
𝑀𝑀𝑐𝑐 𝐻𝐻
𝐵𝐵
+ 𝑃𝑃𝑐𝑐 + ΛCc (1)
𝑃𝑃4 =
𝑃𝑃𝑠𝑠 + cot 𝜔𝜔 + Λ𝐶𝐶𝑠𝑠 (1)

s3' is the minimum support pressure required so that the bearing capacity of the wedge
is higher than the vertical load exerted by the prism and is given by:

𝑠𝑠3′ = 𝐹𝐹1 𝛾𝛾𝑤𝑤 Δℎ + 𝐹𝐹2 𝛾𝛾 ′ 𝐻𝐻 − 𝐹𝐹3 𝑐𝑐 ′


where,
1 𝐻𝐻 tan 𝜔𝜔
𝐹𝐹1 = (𝐶𝐶Δℎ (1) + 𝛼𝛼�
𝐶𝐶𝑠𝑠 (1) 𝐵𝐵

𝑡𝑡∗
−𝜆𝜆 tan 𝜙𝜙′
1 𝐶𝐶𝛾𝛾 (1)𝐵𝐵 tan 𝜔𝜔𝜔𝜔(1 − 𝑒𝑒 𝑅𝑅 )
𝐹𝐹2 = � + �
𝐶𝐶𝑠𝑠 (1) 𝐻𝐻 𝐵𝐵𝐵𝐵 tan 𝜙𝜙 ′

′ 𝑡𝑡
1 tan 𝜔𝜔𝜔𝜔(1 − 𝑒𝑒 −𝜆𝜆 tan 𝜙𝜙 𝑅𝑅 )
𝐹𝐹3 = (𝐶𝐶 (1) +
𝐶𝐶𝑠𝑠 (1) 𝑐𝑐 𝐵𝐵𝐵𝐵 tan 𝜙𝜙 ′

Coefficients used in the equation are:

2λ tan ϕ′
Λ=
cos ω − sin ω tan ϕ′
𝑀𝑀 = 𝑀𝑀𝑐𝑐 𝐵𝐵 2 𝑐𝑐 ′ − 𝑀𝑀𝛾𝛾 𝐵𝐵 3 𝛾𝛾 ′
𝑃𝑃 = 𝑃𝑃𝑐𝑐 𝐵𝐵 2 𝑐𝑐 ′ + 𝑃𝑃𝑠𝑠 𝐵𝐵 2 𝑠𝑠 ′

41
𝑃𝑃𝑠𝑠 = tan(𝜙𝜙 ′ + 𝜔𝜔)

where,
𝑀𝑀𝛾𝛾 = tan 𝜔𝜔
Λ tan 𝜔𝜔
𝑀𝑀𝑐𝑐 =
𝜆𝜆 tan 𝜙𝜙 ′
Λ
𝑃𝑃𝑐𝑐 =
2𝜆𝜆 tan 𝜙𝜙 ′ cos 𝜔𝜔
𝐶𝐶𝜐𝜐 (1) − 1
𝐶𝐶𝑠𝑠 (1) = 𝑃𝑃𝑠𝑠
Λ
𝐹𝐹(1)
𝐶𝐶𝛾𝛾 (1) = 2 𝑀𝑀𝛾𝛾
Λ
𝐶𝐶𝜐𝜐 (1) − 1 𝐹𝐹(1)
𝐶𝐶𝑐𝑐 (1) = 𝑃𝑃𝑐𝑐 + 2 𝑀𝑀𝑐𝑐
Λ Λ
(Λ𝐻𝐻⁄𝐵𝐵 )
𝑐𝑐𝑣𝑣 = 𝑒𝑒
Λ𝐻𝐻
𝐹𝐹(1) = 𝐶𝐶𝑣𝑣 (1) − 1 −
𝐵𝐵
−𝑏𝑏 tan 𝜔𝜔 )
𝑎𝑎(1 − 𝑒𝑒 𝑅𝑅 𝑡𝑡∗ 𝐻𝐻𝐻𝐻 tan 𝜙𝜙′ +𝑅𝑅𝑅𝑅
𝛼𝛼� = (1 − 𝑒𝑒 𝐻𝐻 𝑅𝑅 )
𝑏𝑏 tan 𝜔𝜔 𝐻𝐻𝐻𝐻 tan 𝜙𝜙 ′ + 𝑅𝑅𝑅𝑅

𝑡𝑡 𝑖𝑖𝑖𝑖 𝑐𝑐 ′ ≤ 𝛾𝛾 ′ 𝑅𝑅 𝑜𝑜𝑜𝑜 𝑧𝑧 ∗ ≥ 𝐻𝐻 + 𝑡𝑡;


𝑡𝑡 ∗ = �𝑧𝑧 ∗ − 𝐻𝐻 𝑖𝑖𝑖𝑖 𝑐𝑐 ′ > 𝛾𝛾 ′ 𝑅𝑅 𝑎𝑎𝑎𝑎𝑎𝑎 𝐻𝐻 + 𝑡𝑡 > 𝑧𝑧 ∗ > 𝐻𝐻;
0 𝑖𝑖𝑖𝑖 𝑐𝑐 ′ > 𝛾𝛾 ′ 𝑅𝑅 𝑎𝑎𝑎𝑎𝑎𝑎 𝐻𝐻 > 𝑧𝑧 ∗ ;

Where R is the ratio of the area to the circumference of a horizontal cross-section. The
constants a and b are determined for curve fitting. Value of varies from 4.496 to 2.850
and b varies from 1.935 to 1.638.

The significant improvement over the existing method is that the single wedge method
proposed by Anagnostou and Kovári [2], [7] considers only the average horizontal seepage
force over the height of the wedge and only the resultant normal force (but not the
distribution of the normal stress) at the inclined shear plane. Consequently, the tensile
stresses that may develop locally at the upper part of the wedge remain undetected as
long as the resultant normal force is compressive.

42
Based on the calculations discussed, the authors proposed a dimensionless design
diagrams for necessary support pressure for the condition Overburden/ Equivalent width
= 1. The authors propose that these results are approximately accurate for higher
overburden also. Further authors produced nomograms for two condition: γw/γ’ = 1.0
and for γw/γ’ = 0.8. Design charts for the conservative case (γw/γ’ = 1.0) for various head
difference are produced below for reference.

Figure 41. Design diagrams for effective support pressure, ∆h/H = 0-2 [10]

43
Figure 42. Design diagrams for effective support pressure, ∆h/H = 3-5 [10]

2.2 Comparison of different models

2.2.1 Cohesive Soils

Most of the models for soft cohesive soils are derived for C/D less than 5 and hence our
comparison of different models are restricted to that ratio. Face pressure values are
compared for a surcharge value of 20 kPa in a soil with unit weight γ = 20 kN/m3 and
cu = 50 kPa. The results of the comparison are shown below in Figure 43.

44
450

400

350
Face Pressure in kPa

300

250

200

150

100

50

0
0 1 2 3 4 5 6
C/D Ratio
Broms and Bennermark method Davis method - Cylindrical
Davis method - Spherical Kimura & Mair
Mori - Blowout Augarde et al - Lower
Klar et al

Figure 43. Comparison of face pressure for different cohesive soil model

The results indicate that face pressure calculated using Davis method result in a
conservative face pressure values compared to the recent model by Klar et al 2007 [23].
The face pressure using Mori's model are high because it relates to blowout condition.
To check the sensitivity of undrained shear strength on the estimated face pressure
values, a comparison is made by changing the cu values while retaining C/D ratio of 2
(Overburden = 10m, Diameter = 5m) with a surcharge of 20 kPa. Unit weight and
undrained shear strength are considered constant with depth. Comparison are presented
in Figure 44. It is observed that face pressure estimation by Davis method [11] is higher
as noted in previous plot. Compared to other methods shown in the plot, Davis method
[11] is less sensitive to increase in the undrained shear strength. Minimum face pressure
estimated by Augarde et al [22] is underestimated for lower undrained shear strength
values.

45
300

250

200
Face Pressure in kPa

150

100

50

0
0 20 40 60 80 100 120
Undrained Shear Stength Cu [kPa]
Davis method - Spherical Kimura & Mair
Ellstein Augarde et al - Upper
Klar et al
Figure 44. Comparison of face pressure for different Cu values

2.2.2 Cohesionless Soils

Minimum face stability is compared for Cohesionless soil and presented for different C/D
values. For the case of comparison, even Mohr-Coulomb models are considered but with
very less drained cohesion (5 kPa). Surcharge (20 kPa), unit weight of soil (20 kN/m3)
and ϕ' (30°) are kept constant for this comparison. The results are plotted in Figure 45.
The face pressure values using Atkinson and Potts method [26] are very conservative
and are beyond the extents of this plot. An evident observation in the comparison is
that the face pressure for cohesionless soils are independent of the overburden.
Anagnostou & Kovari's method is used without considering the seepage forces.

46
45

40

35
Face Pressure in kPa

30

25

20

15

10

0
0 1 2 3 4 5
C/D Ratio
Krause - Quarter Circle (c' = 5) Leca & Dormieux - MII (c' = 5)

Caquot/ C.Carranza-Torres (c' = 5) Kolymbas Method (c' = 5)

Anagnostou 2012 (c' = 5) Chen, Tang, Yinm et al

Vermeer' Empirical Method (c' = 5) Mollon (c' = 5)

Figure 45. Comparison of face pressure for Cohesionless soils

It can be observed that when the relative depth C/D is less than 1, there is a slight
increase in support pressure with the overburden. Above C/D of 1, there is no change
in the minimum support pressure. Support pressure calculated using Mollon's method
[14], Leca & Dormieus method [13], Anagnostou's method [9], Vermeer's empirical [29]
method and Anagnostou & Kovari's method [6] are comparable. Face pressure values
calculated using other methods are conservative. Face pressure values calculated using
Jancsecz and Steiner [33] are linearly increasing with the overburden as it is based on
Terzaghi's load theory for shallow tunnels.

47
200

180

160

140
Face Pressure in kPa

120

100

80

60

40

20

0
15 20 25 30 35 40 45 50
Phi [°]
Krause - Quarter Circle (c' = 5)
Leca & Dormieux - MII (c' = 5)
Caquot/ C.Carranza-Torres (c' = 5)
Kolymbas Method (c' = 5)
Anagnostou 2012 (c' = 5)
Chen, Tang, Yinm et al
Vermeer' Empirical Method (c' = 5)
Mollon (c' = 5)
Atkinson and Potts
Jancsecz & Steiner (c' = 5)
Anagnostou & Kovari 1996 (c' = 5, no seepage)

Figure 46. Comparison of face pressure for different ϕ values

To understand the influence of ϕ on the minimum face pressure, ϕ' angle is varied for
the above models for the condition C/D = 1.5 (Overburden = 7.5m and Tunnel Diameter
= 5.0m), with surcharge of 20 kPa and unit weight of 20kN/m3. For simplicity, ground
water is not considered and hence no seepage forces are included. The results are
presented in Figure 46. It can be observed that the face pressure estimates by latest

48
models (Anagnostou [9], Chen et al [28] and Mollon [14]) are closer. Face pressure
estimation by Caquot's model [25] are most sensitive to friction angle.

2.2.3 C-Phi Soil

Similar to the behaviour in cohesionless soil, even face pressure in C-Phi soil stay
constant or mildly influenced with the change in overburden.

160

140

120
Face Pressure in kPa

100

80

60

40

20

0
0 1 2 3 4 5
C/D Ratio
Jancsecz & Steiner Krause - Quarter Circle

Anagnostou & Kovari 1996 (no seepage) Caquot/ C.Carranza-Torres

Kolymbas Method Vermeer' Empirical Method

Anagnostou 2012 Mollon 2009

Figure 47. Comparison of face pressure for c-ϕ soil

Figure 47 presents the comparison of face pressure values using models available for c-ϕ
soils. Anagnostou and Kovari's method [6] is compared without considering the seepage
forces. The face pressure estimated in Figure 47 are based on effective cohesion of 25
kPa and ϕ' of 20°. Surcharge and unit weight of soil is considered at 50 kPa and 20

49
kN/m3. Face pressure values estimated by Anagnostou 2012 [9] is constant with the
change in C/D as it is independent of overburden. Anagnostou and Kovari's method [6]
is slightly dependent on overburden depth. It can also be observed that face pressure
estimated using this method are conservative in nature.
700

600

500
Face Pressure in kPa

400

300

200

100

0
20 30 40 50 60 70 80 90 100
c' [kPa]
Jancsecz & Steiner Krause - Quarter Circle
Anagnostou & Kovari 1996 (no seepage) Caquot/ C.Carranza-Torres - Spherical
Kolymbas Method Vermeer' Empirical Method
Anagnostou 2012 Mollon 2009

Figure 48. Comparison of face pressure for c-ϕ soil - influence of cohesion

Figure 48 compares the influence of value of cohesion on the face pressure estimation. It
can be observed that most of the models have equal effect on the change of cohesion.
Kolumbas' method [35] seems to be less sensitive to the change in cohesion. Its again
noticed that the face pressure values estimated using Anagnostou 2012 [9], and Mollon
[14] are comparably similar. Face pressure estimation by Jancsecz's method [33] is
independent of cohesion and the values estimated are beyond the bounds of the graph,
hence it is not shown in the figure.

50
Above comparison is made for a C/D ratio of 1.5m (Overburden = 15m and Diameter
= 10m). Constant surcharge (50 kPa), soil unit weight (20 kN/m3) and ϕ (20°) is
considered for this comparison.

2.2.4 Models which consider Seepage forces

The effect of seepage flow on the face stability was reported by Anagnostou and Kovari
[2], [7] and presented nomograms for the assessment of the required effective support
pressure for different hydraulic boundary conditions. Later, Lee et al. [38] considered
seepage forces in the upper bound solution of Leca and Dormieux [13]. While Lee et al.
[38] investigated only the case of purely frictional soils, Park et. al [37] investigated the
same in a cohesive frictional soil. It was only recently, Perazzelliet. al 2014 [10] presented
dimensionless design normalized diagrams (shown in Figure 41 and Figure 42) for
computation of effective support pressure using method of slices for seepage conditions.
In this section, face pressure values are calculated considering seepage pressure using
two different models - Anagnostou and Kovari 1996 [2] and Perazzelli et. al 2014 [10]
are compared and discussed.

Effective face pressure is estimated for 10m dia tunnel with the condition: Overburden
/ Diameter = 1, c’ = 0 and hydraulic head ∆h = 30m for various ϕ’ values and presented
in Figure 49.

51
350

300
Effective Face Pressure in kPa

250

200

150

100

50

0
15 17.5 20 22.5 25 27.5 30 32.5 35
ϕ' [°]
Perazzelli, Leone, Anagnostou 2014 Anagnostou & Kovari 1996

Figure 49. Comparison of effective face pressure as a function of ϕ’

It is observed that the effective face pressure estimated using Perazzelli et al nomograms
are constantly lower than that of the Anagnostou 1996. Constant difference is
maintained even with the increase of ϕ’ values. This suggests that the method of slices
leads to lower effective support pressure values (for equilibrium condition). The same
comparison is repeated with a constant ϕ’ (= 25°) but now varying the effective cohesion
instead, and results are presented in Figure 50. However, face pressure calculated using
Anagnostou 1996’s nomograms are considerably lower than the one Perazzelli et al’s
nomograms for higher c’ values. This is because, Anagnostou et al 1996 [2] considers
only equilibrium of the prism and does not check the tensile failure.

52
250

200
Effective Face Pressure in kPa

150

100

50

0
0 20 40 60 80 100 120
c' [kPa]
Perazzelli, Leone, Anagnostou 2014 Anagnostou & Kovari 1996

Figure 50. Comparison of effective face pressure as a function of c’

Thus, in case of high hydraulic gradient and if the cohesion of the ground is high (which
may be true for weak rocks), the necessary effective face support pressure may be much
higher than the pressure required for the stability of the wedge. Because, in this case,
tensile failure rather than sliding becomes the critical mode for the determination of
support pressure [10]. This means that, in such situations nomograms of Anagnostou et
al. 1996 [2] may underestimate the necessary support pressure and thus may be unsafe.
This effect is further studied by comparing effective face pressure with varying hydraulic
gradient for two different cohesion value (0 and 100 kPa), using both the methods and
is presented in Figure 51. Results indicate that, as observed above, the results from
Anagnostou et al 1996 [2] are underestimating the support pressure at higher cohesion.
Another important observation from Figure 51 is, as the hydraulic gradient increases,
the estimate by Anagnostou et al 1996 [2] is approaching the values estimated using
Perazzelli et al. (i.e, the governing mechanism is changing from tensile failure back to
limit equilibrium failure).

53
600

500
Effective Face Pressure in kPa

400

300

200

100

For a constant ϕ' = 25°


0
0 1 2 3 4 5 6 7 8 9 10
∆h/H
Perazzelli, Leone, Anagnostou 2014, c' =0
Anagnostou & Kovari 1996, c'=0
Perazzelli, Leone, Anagnostou 2014, c' = 100 kPa
Anagnostou & Kovari 1996, c' = 100 kPa

Figure 51. Comparison of face pressure as a function of hydraulic gradient

2.2.5 Models for Blowout Pressure

Two models to estimate blowout pressure are discussed in this study. One of them is for
Soft cohesionless soil (Mori 1991 [20]) and the second model is for C-Phi soils (MIII
mechanism of Leca and Dormieux [13]). In order to compare, the blowout pressure is
estimated for a shallow tunnel (C/D of 1, with C = 5m) using both the methods for
different values of undrained cohesion. ϕ value of 1 degree is used in Leca and Dormieux's
model. A constant surcharge and unit weight of 20kPa and 20kN/m3 is considered.

54
700

600
Blow Out Pressure in kPa

500

400

300

200

100

0
0 10 20 30 40 50
Cu [kPa]

Leca & Dormieux - MIII (Phi = 1 degree) - 1990 Mori - 1991

Figure 52. Comparison of Blowout pressure

From Figure 52 it is evident that blowout pressure estimation using Leca and Dormieux
[13] method is very sensitive to the Cu value. Mori's method [20] is based on fracturing
model in cohesive soils whereas Leca's model is primarily derived for frictional soil and
then extended to general C-Phi soil. This can be the reason for wide difference in the
blowout pressure estimation at higher Cu value. Hence model selection for a particular
site needs to selected with due care.

2.3 Case Study

In order to apply the theory discussed so far, a real life case study is presented in this
section. Project under consideration is part of a Mass Rapid Transit (MRT) metro
project. The project includes twin bored tunnels with outer diameter of 6.35m. The bored
tunnels are being constructed using an Earth Pressure Balanced (EPB) Tunnel Boring
Machine (TBM). Pre-cast concrete segmental lining is installed within the tail shield of

55
the TBM, designed to provide final support for the tunnel. The length of the tunnel
considered for this case study is 0.7km between two metro stations.
The TBM is excavated to encounter weathered Granite and soil during its drive. The
soil profile typically consists of fill followed by fluvial sands, fluvial clays, estuarine and
Granite in its various degree of decomposition. The soil profile along the tunnel
alignment is shown in Figure 53. Dark line represents the interface between the Granite
and soil. Based on the geological profile, it can be observed that there are many mixed
face conditions expected along the tunnel drive.

Figure 53. Geological profile and tunnel alignment – Case study

TBM operating pressure is estimated using the theory discussed in section 2.1. Stability
model by Anagnostou et al 1996 [2] and their nomograms are used to calculate effective
support pressure while considering the seepage forces. The total face support pressure is
the sum of effective face support pressure and relevant pore water pressure in the TBM
cutter head chamber.
The imposed loads comprise of surface surcharge of 20 kPa and gravity loading of soil
above the sliding wedge. Ground water table at 1m below the ground surface is

56
considered in calculations. Both drained and undrained behaviour of relevant soil
materials are examined.
TBM face support pressure required to maintain stable face conditions at the cutter
head during TBM driving has been performed at analysis sections based on every 10m
interval. A total of 70 number of analysis sections have been calculated for the alignment
shown in Figure 53. Both serviceability (SLS) and ultimate (ULS) limit states have been
considered. For ULS calculations, minimum partial factors have been applied to soil
parameters -ϕ’ and c’ divided by 1.25 (for ϕ’, the reduction is applied to tanϕ’).
It is observed that when the tunnel is passing through ground consisting of interbedded
sands and clays, it is not possible, using the simple charts, to account for the combination
of the different types of soil. Averaging the parameters with different weights for crown,
upper face, invert etc could be done to reach a pseudo-realistic behaviour. However, use
of the Horn model, or its derivatives, is not appropriate where there are substantial
thicknesses of clay, as the failure surface in clay does not correspond to that used for the
Horn model.
This limitation is overcome by using separate analyses for the beds of clay or sand in
and over the face of the tunnel. The highest calculated face pressure can then be adopted
as the Target / Recommended pressure. At this point, understanding the limitations /
behaviour of different stability models gives a better judgement in selecting appropriate
stability models. This approach should be conservative, in relation to the actual
conditions. Using it, a variety of possible ground conditions are allowed for, giving a
significant degree of robustness to the calculations. Based on these considerations, the
minimum total face pressure, Recommended / Target face pressure and maximum face
pressure is calculated and presented in Figure 54. In addition, other relevant details like
hydrostatic water pressure and octahedral stress are also presented in the chart.

57
Figure 54. Estimated TBM operating face pressure – Case study

58
3 Probability Analysis

The most widely accepted and successful way to deal with the uncertainties inherent in
dealing with geological materials came to be known as the observational method. It is
also an essential part of the New Austrian Tunnelling Method [41]. The observational
method grew out of the fact that it is not feasible in many geotechnical applications to
assume very conservative values of the loads and material properties and design for those
conditions. The resulting design is often physically or financially impossible to build.
Instead the engineer makes reasonable estimates of the parameters and of the amounts
by which they could deviate from the expected values. Then the design is based on
expected values - or on some conservative but feasible extension of the expected values
- but provision is made for action to deal with the occurrence of loads or resistances that
fall outside the design range. During construction and operation of the facility,
observations of its performance are made so that appropriate corrective action can be
made. This is not simply a matter of designing for an expected set of conditions and
doing something to fix any troubles that arise. It involves considering the effects of the
possible range of values of the parameters and having in place a plan to deal with
occurrences that fall outside the expected range. It requires the ongoing involvement of
the designers during the construction and operation of the facility.
As discussed above, one of the traditional ways used to address uncertainty is to use
factor of safety design approach. More recent alternatives are the load and resistance
factor design (LRFD) approach and the characteristic values and partial factors used in
the limit state design approach in Eurocode 7. Yet another approach can play at least a
useful complementary role to LRFD and Eurocode 7, namely the design based on a
target reliability index that explicitly reflects the uncertainty of the parameters and their
correlation structure. It must be emphasized at the outset that reliability approaches do
not remove uncertainty and do not alleviate the need for judgment in dealing with the
world. They do provide a way of quantifying those uncertainties and handling them
consistently [42].

59
The simplest definition of the term reliability is the probability of a desirable and
sustainable function of a system. Engineering reliability, as in mathematics, is a basic
science which may be used in tunnelling as well. Since years ago, many researchers have
been dealing with the issue of reliability of structures. Also, some researchers have tended
to work on reliability in geotechnical structures. Various methods for application of
reliability analysis in engineering problems can be found out in Dai and Wang [43]. There
are also some references that indicate the application of reliability engineering specifically
on geotechnical engineering.

Reliability analyses, involving neither complex theory nor unfamiliar terms, can be used
in routine geotechnical engineering practice. These reliability analyses require little effort
beyond that involved in conventional geotechnical analyses. Baecher and Christian [42]
reported many fundamental approaches for reliability practices in geotechnical field.
More recently, Duncan [44] did a comprehensive study on the factors of safety and
reliability in geotechnical engineering. Reliability evaluation can provide a means of
evaluating the combined effects of uncertainties in the parameters involved in the
calculations, and offer a useful supplement to conventional structural analyses.
Engineering reliability is based on the principles of probability and mathematics which
is now a useful tool for estimating the proper function of systems. Reliability based
design calls for a willingness to accept the fundamental philosophy that
• Absolute reliability is an unattainable goal in the presence of uncertainty
• Probability theory can provide a formal framework for developing design criteria
that would ensure that probability of failure is acceptably small.
This also provides a consistent method for propagation of uncertainties and a unifying
framework for risk assessment across disciplines and national boundaries [45].

60
Figure 55. General stochastic problem [45]

3.1 Uncertainties

Two main sources of geotechnical uncertainties can be distinguished. The first arises
from the evaluation of design soil properties, such as undrained shear strength and
effective stress friction angle. This source of geotechnical uncertainty is complex and
depends on inherent soil variability, degree of equipment and procedural control
maintained during site investigation, and precision of the correlation model used to relate
field measurement with design soil property. Based on Phoon [45] three ranges of soil
property variability (low, medium, high) were found to be sufficient to achieve
reasonably uniform reliability levels for simplified reliability based checks.

Figure 56. COV of geotechnical parameters [45]

The second source arises from geotechnical calculation models. Although many
geotechnical calculation models are “simple,” reasonable predictions of fairly complex

61
soil–structure interaction behaviour still can be achieved through empirical calibrations.
Model factors, defined as the ratio of the measured response to the calculated response,
usually are used to correct or simplifications in the calculation models.

It is evident that a geotechnical parameter (soil property or model factor) exhibiting a


range of values, possibly occurring at unequal frequencies, is best modelled as a random
variable. The existing practice of selecting one characteristic value (e.g. mean, “cautious”
estimate, 5% exclusion limit, etc.) is attractive for designers, because design calculations
can be carried out easily using only one set of input values once they are selected.
However, this simplicity could be deceptive [45]. The choice of the characteristic values
clearly affects the overall safety of the design, but there are no simple means of ensuring
that the selected values will achieve a consistent level of safety.

3.2 Monte Carlo Simulation

A wide range of engineering and scientific disciplines use simulation methods based on
randomized input, often called Monte Carlo methods. They have been employed to study
both stochastic and deterministic systems. Because developments are often bound to the
application disciplines, notation and nomenclature sometimes reflect the preferences of
a particular field. Monte Carlo methods can be divided into two broad areas. First, there
is the simulation of a process that is fundamentally stochastic. The second area involves
problems that are not inherently stochastic but can be solved by simulation with random
variables. Typically, Monte Carlo method require that the function be evaluated a large
number of points. This becomes easy with the present computing systems. The idea of
Monte Carlo methods is the generation of random events in a computer model and this
generation is repeated many times and count the occurrence number of specific
conditions. The basic procedure of Monte Carlo method is:
• Define a domain of possible events
• Generate events randomly
• Perform deterministic judgements of system based on the events

62
• Count the occurrence number of a specific system state among total observations

Having assigned PDF for input parameters considered as random variables, the
probabilistic assessment of the geotechnical parameters is accomplished. The Monte
Carlo simulation is applied to obtain minimum face stability pressure using the Microsoft
Excel program. In the Monte Carlo simulation the values of each input parameter is
generated randomly considering the PDF.
The advantage of the Monte Carlo technique compared to other probabilistic methods
is the possibility of getting complete PDF of output. The main disadvantage of the
method is the large number of simulations to be executed.

3.3 Probability Assessment

In a probabilistic approach the input parameters can be deterministic and probabilistic.


Deterministic parameters are considered as fixed parameters with single values. In case
of probabilistic parameter, a probability density function (PDF) characterized by two
statistical parameters namely mean and coefficient of variation (COV) has to be
assigned. Normally PDF and its statistical parameter should be assigned by analysing
the laboratory or field test results and/or measured data. Nevertheless for the parameters
whose data are insufficient to fit a PDF known distributions for such parameter are
assumed / normal distribution is considered.

The problem can be idealized by considering a circular rigid tunnel of diameter D driven
under a depth of cover C. A surcharge qs is applied at the ground surface
Due to uncertainties in the soil shear strength parameters, the cohesion c, and the angle
of internal friction ϕ are considered as random variables. Probability assessment is
carried out for three models - Mollon et al 2010 [14], Anagnostou 2012[9] and Chen et al
2014 [28]. Results of the Monte Carlo simulation are presented in this section. Monte
Carlo simulation is performed using simple spreadsheet software. The pseudo random
number generator is used to simulate failure times of individual units in the system.

63
Logical expressions are then used to determine system success or failure. (PDF) and
cumulative distribution function (CDF) of the support pressure are presented.

In order to explain the procedure used for Monte Carlo (MC) simulation, the work flow
is explained using a simple model at first and then outputs using other models are
presented and compared.

Figure 57 shows the input required for performing Monte Carlo simulation for a simple
closed form model using Anagnostou, 2012 [9]. In addition to the four variables in the
equation, additional inputs to indicate the expected variation in the parameters are
required. This could be defined using statistical parameters, mean and coefficient of
variation. However, for simplicity, only absolute maximum and minimum values of each
parameters are used as input (the random generator considers equal probability for each
value in that range). No variation is considered in diameter.

Figure 57. Input values for MC simulation using Anagnostou, 2012

Using the parameters as discussed above, Monte Carlo simulation is run with a sample
size of 5000. For each computation, the variables are randomly selected within the
bounds of expected variation.

64
Figure 58. Snapshot of Random number generation

Figure 59 presents the histogram of one such iteration. More number of iterations could
be performed to refine the results. It could be observed from the histogram plot that
with the expected variation in the input paramters, the most likely estimate of support
pressure (using Anagnostou, 2012 [9]) is around 42 kPa. However, there are chances that
the expected variablity in the parametes could result in support pressure as high as 135
kPa.

65
600

500

400
Count

300

200

100

0
17.7
25.6
33.4
41.3
49.2
57.1
65.0
72.9
80.7
88.6
96.5
1.9
9.8

104.4
112.3
120.2
128.0
135.9
143.8
151.7
159.6
Support Pressure [kPa]
Figure 59. Histogram of MC simulation using Anagnostou, 2012

1.00
0.90
0.80
Cumulative Probablity

0.70
0.60
0.50
0.40
0.30
0.20
0.10
0.00
104
112
120
128
136
144
151
159
10
17
25
33
41
49
57
65
73
81
88
96
2

Face Pressure [kPa]

Figure 60. CDF of MC simulation using Anagnostou, 2012

66
Figure 61. Summary of MC simulation using Anagnostou, 2012 in Excel

Since the simulation generates the PDF and CDF of the required support pressure, we
can estimate the minimum support pressure for the required reliability percentage
instead of (over-conservatively) assuming certain factor of safety. Here in this case, the
factor of safety that would have resulted in 95 percentile to result is back calculated and
found to be around 1.8.

By varying the input parameters, it is observed that, for the required confidence on the
minimum support pressure, (back calculated) factor of safety varies with the physical
configuration and spread of input parameters. Hence, the minimum support pressure
calculated using probabilistic analysis are more reliable and relevant to site conditions /
uncertainties.

Similar probabilistic assessment is performed for cohesionless soils using Chen et al [28]
model (by varying γ and ϕ as assumed in Figure 57, with c = 0 and C/D = 1.5) and the

67
results are presented in Figure 62. It is observed that results using this model have higher
central tendency and follow an ideal normal distribution. It is to be noted that for such
probability distribution, higher factor of safety values are not required even for most
reliable estimate of face pressure.

700 1.00

0.90
600
0.80

Cumulative Probability
500 0.70

0.60
400
Count

0.50
300
0.40

200 0.30

0.20
100
0.10

0 0.00
13
15
16
18
19
21
22
24
25
27
28
30
31
33
34
36
37
39
40
41
43

Support pressure [kPa]

Figure 62. Output of MC simulation using Chen et al 2014 model


Same exercise is repeated for a c-ϕ soil using Mollon et al 2010 [14] model (by varying
c, γ and ϕ as assumed in Figure 57 and C/D = 1.5) and the results are presented in
Figure 63. It is observed that the results also possess similar central tendency compared
to Chen et al model (Figure 62) but has a higher standard deviation. However Skewness
and Kurtosis of both the probability distribution function are comparable.

68
350 1.00

0.90
300
0.80

Cumulative Probability
250 0.70

0.60
200
Count

0.50
150
0.40

100 0.30

0.20
50
0.10

0 0.00
11
14
17
19
22
25
28
31
34
37
40
43
46
49
52
55
58
2
5
8
-1

Support Pressure [kPa]


Figure 63. Output of MC simulation using Mollon et al 2010 model

Based on this procedure, Monte Carlo simulations are performed for Face stability
models which consider seepage forces (Anagnostou and Kovari 1996 [2] and Perazzelli
et. al 2014 [10] ) and important observations are noted.

For this analysis, effective cohesion c’ and effective friction angle ϕ’ are assumed to
follow normal distribution. Parameters for c’ are: mean = 50 kPa, standard deviation =
20 kPa. Parameters for ϕ’ are: mean = 25°, standard deviation = 5°. The distribution
of c’ and ϕ’ considered in the analysis is presented in Figure 64. C/D ratio of 1 is
considered with a constant hydraulic gradient, ∆h/H of 3. Submerged unit weight of soil
is taken as 10 kN/m3.

69
.
Figure 64. Distribution of input parameters for MC simulation

Monte Carlo simulation is run with a sample size of 10,000 for the both the methods
and Probability Distribution Function of the effective face pressure are compared in
Figure 65. Green histogram are corresponding to the results from Anagnostou and Kovari
1996 and Red histogram are the values corresponding to Perazzelli et. al 2014 [10].

0.016

0.014

0.012

0.01
Frequency

0.008

0.006

0.004

0.002

0
-200

-150

-100

100

150

200

250

300

350
-50

50
0

Effective face pressure [kPa]

Figure 65. Comparison of PDF using the two methods

Some of the important observations from this comparison are, effective face pressure
values estimated using Anagnostou and Kovari 1996 [2] have wider spread of possible

70
values. That means, higher factor of safety needs to be used using this method to account
for various uncertainties. Whereas, the estimation by the second method leads to values
with higher central tendency and factor of safety required is relatively less to achieve
the same reliability. In summary, a logical method to select the appropriate model for
face pressure estimation is discussed using simple analytical tools. More importantly, an
approach to consider the uncertainties in the design (rather than relying on over
conservative factor of safety) is discussed.

71
4 Results and Conclusion

In preparation of this report, the author has investigated different available models for
face pressure calculation for different types of soils. Recent development of stability
models were discussed and compared with widely used methods in the industry. With
the constant process of ‘pushing the envelope’, bigger, shallower (or deeper) tunnels, and
ever more stringent criteria, expansion of the existing design nomograms / curves, to
cover lower and higher C/D ratios than currently available, would also be useful.

Based on the sensitivity analysis presented in section 2.2 and the case study presented
in section 2.3, we can deduce that it is very important to understand the limitations of
each model. Although there has been efforts to formulate a generalized model for all
types of soils following Mohr-Coulomb criterion, some models are found to be sensitive
for certain range of input values.

A probabilistic model based on simple spreadsheet is proposed to calculate the unbiased


limiting tunnel collapse pressure. Explanations have been given for a Monte Carlo
simulation method for a reliability analysis. Detailed explanations have been given for
the application of Excel software to Monte Carlo simulation. Monte Carlo method can
straightforwardly treat any kind of system structure and operating conditions. We
therefore can avoid the restrictive modelling assumptions that had to be introduced in
analytical methods and even use empirical methods (if required). It should be noted that
all of the random parameters were assumed to be independent for simplicity. This could
be further studied by considering the dependent variables. Using this simulation, the
probability of failure for a specified applied pressure at the tunnel face could be estimated
or conversely the minimum face pressure for the required degree of reliability could be
estimated. This approach leads to more confident judgement of the face pressure than
using single Factor of Safety for all the cases.

Similar to the closed faced TBMs, this approach could be used in conventional tunnelling
/ SEM / NATM to determine probability of face failure / probability of tunnel face

72
deformation using different support mechanism and assess related risks or select most
reliable support mechanism.

Although the calculations are an essential part in the development of suitable operating
pressures, a significant degree of engineering judgement needs to be applied both before
and after the calculations. Before carrying out the calculations the ground and
groundwater model has to be developed, select critical sections etc. and after
calculations, we need to assess the risks of encountering more adverse conditions
(requiring significantly higher pressure), understand the influence of interfaces etc.

73
Bibliography
[1] V. Guglielmetti, P. Grasso, A. Mahtab, and S. Xu, Mechanized tunnelling in urban
areas: design methodology and construction control. CRC Press, 2008.

[2] G. Anagnostou and K. Kovári, “Face stability conditions with earth-pressure-


balanced shields,” Tunnelling and Underground Space Technology, vol. 11, no. 2,
pp. 165–173, Apr. 1996.

[3] X.-W. Tang, W. Liu, B. Albers, and S. Savidis, “Upper bound analysis of tunnel
face stability in layered soils,” Acta Geotechnica, vol. 9, no. 4, pp. 661–671, Jul.
2014.

[4] R. Chen, J. Li, L. Kong, and L. Tang, “Experimental study on face instability of
shield tunnel in sand,” Tunnelling and Underground Space Technology, vol. 33,
pp. 12–21, Jan. 2013.

[5] N. Horn, “Horizontaler Erddruck auf senkrechte Abschlussflaechen von


Tunnelroehren,” Landeskonferenz der Ung- arischen Tiefbauindustrie, pp. 7–16,
1961.

[6] G. Anagnostou and K. Kovári, “Face stability in slurry and EPB shield
tunnelling,” in Geotechnical Aspects of Underground Construction in Soft Ground,
1996, pp. 453–458.

[7] G. Anagnostou and K. Kovári, “The face stability of slurry-shield-driven tunnels,”


Tunnelling and Underground Space Technology, vol. 9, no. 2, pp. 165–174, Apr.
1994.

[8] W. Broere, “Tunnel Face Stability & New CPT Applications,” TU Delft, 2001.

[9] G. Anagnostou, “The contribution of horizontal arching to tunnel face stability,”


geotechnik, vol. 35, no. 1, pp. 34–44, 2012.

[10] P. Perazzelli, T. Leone, and G. Anagnostou, “Tunnel face stability under seepage
flow conditions,” Tunnelling and Underground Space Technology, vol. 43, pp. 459–
469, Jul. 2014.

74
[11] E. H. Davis, R. J. Mair, H. N. Seneviratine, and M. J. Gunn, “The stability of
shallow tunnels and underground openings in cohesive material,” Géotechnique,
vol. 30, no. 4, pp. 397–416, 1980.

[12] H.-B. Mühlhaus, “Lower bound solutions for circular tunnels in two and three
dimensions,” Rock Mechanics and Rock Engineering, vol. 18, no. 1, pp. 37–52,
1985.

[13] E. Leca and L. Dormieux, “Upper and lower bound solutions for the face stability
of shallow circular tunnels in frictional material,” Géotechnique, vol. 40, no. 4, pp.
581–606, 1990.

[14] G. Mollon, D. Dias, and A. Soubra, “Face Stability Analysis of Circular Tunnels
Driven by a Pressurized Shield,” no. January, pp. 215–229, 2010.

[15] J. H. Atkinson and R. J. Mair, “Soil mechanics aspects of soft ground tunnelling,”
Ground Engineering, vol. 14, no. 5, pp. 20–38, 1981.

[16] R. J. Mair and R. N. Taylor, “Theme lecture: Bored tunneling in the urban
environment,” of XIV ICSMFE [131], pp. 2353–2385, 1999.

[17] B. B. Broms and H. Bennermark, “Stability of clay at vertical openings,” ASCE


Journal of the Soil Mechanics and Foundations Division, vol. 93, no. 1, pp. 71–
94, 1967.

[18] T. Kimura and R. J. Mair, “Centrifugal testing of model tunnels in soft clay,” in
Proceedings of the 10th International Conference on Soil Mechanics and
Foundation Engineering, 1981, pp. 319–322.

[19] A. R. Ellstein, “Heading failure of lined tunnels in soft soil,” Tunnels and
Tunnelling, pp. 51–54, 1986.

[20] A. Mori, M. Tamura, K. Kurihara, and H. Shibata, “A suitable slurry pressure in


slurry-type shield tunneling,” in Tunneling ’91, Institution of Mining and
Metallurgy, 1991, pp. 361–369.

[21] A. Bezuijen, M. Huisman, and D. R. Mastbergen, “Verdringingsprocessen bij


gestuurd boren-Technical Report 17,” in Boren Tunnels en Leidingen, 1996.

75
[22] C. E. Augarde, A. V. Lyamin, and S. W. Sloan, “Stability of an undrained plane
strain heading revisited,” Computers and Geotechnics, vol. 30, no. 5, pp. 419–430,
Jul. 2003.

[23] A. Klar, A. S. Osman, and M. Bolton, “2D and 3D upper bound solutions for
tunnel excavation using ‘elastic’ flow fields,” International Journal for Numerical
and Analytical Methods in Geomechanics, vol. 31, no. 12, pp. 1367–1374, 2007.

[24] H. Tresca, “On the Flow of Solid Bodies Subjected to High Pressures,” Comptes
Rendus de l’Académie des Sciences, vol. 59, p. 754, 1864.

[25] C. Carranza-Torres, T. Reich, and D. Saftner, “Stability of shallow circular


tunnels in soils using analytical and numerical models,” in 61st Minnesota Annual
Geotechnical Engineering Conference . University of Minnesota, 2013.

[26] J. H. Atkinson and D. M. Potts, “Stability of a shallow circular tunnel in


cohesionless soil,” Geotechnique, vol. 27, no. 2, pp. 203–215, 1977.

[27] P. Chambon and J. Corté, “Shallow Tunnels in Cohesionless Soil: Stability of


Tunnel Face,” Journal of Geotechnical Engineering, vol. 120, no. 7, pp. 1148–
1165, 1994.

[28] R. P. Chen, L. J. Tang, X. S. Yin, Y. M. Chen, and X. C. Bian, “An improved


3D wedge-prism model for the face stability analysis of the shield tunnel in
cohesionless soils,” Acta Geotechnica, pp. 1–10, Feb. 2014.

[29] P. A. Vermeer, N. Ruse, and T. Marcher, “Tunnel heading stability in drained


ground,” Felsbau, vol. 20, no. 6, pp. 8–18, 2002.

[30] S. Kanayasu, I. Kubota, and N. Shikibu, “Stability of face during shield tunneling
– A survey of Japanese shield tunneling,” Fujita and Kusakabe, vol. 71, pp. 337–
343, 1966.

[31] T. Krause, “Schildvortrieb mit flüssigkeits und erdgestützter Ortsbrust,” PhD


Thesis, Technischen Universität Carolo-Wilhelmina, Braunschweig, 1987.

[32] M. Mohkam and Y. . Wong, “Three dimensional stability analysis of the tunnel
face under fluid pressure,” in Numerical Methods in Geomechanics, 1989, pp.
2271–2278.

76
[33] S. Jancsecz and W. Steiner, “Face support for a large mix-shield in heterogeneous
ground conditions,” in Tunnelling’94, Springer, 1994, pp. 531–550.

[34] P. A. Vermeer and N. Ruse, “Die Stabilitat der Tunnelortsbrust im homogenen


Baugrund,” Geotechnik, vol. 24, no. 3, pp. 186–193, 2001.

[35] A. Kirsch and D. Kolymbas, “Theoretische Untersuchung zur Ortsbruststabilität,”


Bautechnik, vol. 82, no. 7, pp. 449–456, 2005.

[36] D. Kolymbas, Tunnelling and tunnel mechanics: A rational approach to tunnelling.


Springer, 2005.

[37] J. K. Park, J. T. Blackburn, and J. Ahn, “Upper bound solutions for tunnel face
stability considering seepage and strength increase with depth,” in Underground
Space – the 4th Dimension of Metropolises Barták, Hrdina, Romancov & Zlámal
(eds), 2007, no. c, pp. 1217–1222.

[38] I. M. Lee and S. W. Nam, “The study of seepage forces acting on the tunnel lining
and tunnel face in shallow tunnels,” Tunnelling and Underground Space
Technology, vol. 16, no. 1, pp. 31–40, 2001.

[39] B. P. A. Vermeer, N. Ruse, and T. Marcher, “Tunnel Heading Stability in Drained


Ground,” vol. 20, no. 6, 2002.

[40] P. Perazzelli, T. Leone, and G. Anagnostou, “A limit equilibrium method for the
assessment of the tunnel face stability taking into account seepage forces,” in
World Tunnel Conference 2013 Geneva, 2013, pp. 715–722.

[41] L. . Rabcewicz, “The New Austrian Tunnelling Method (Part 1 to 3),” Water
Power, 1964.

[42] G. B. Baecher and J. T. Christian, Reliability and Statistics in Geotechnical


Engineering. 2003.

[43] S.-H. Dai and M.-O. Wang, Reliability analysis in engineering applications. Van
Nostrand Reinhold, 1992.

[44] J. M. Duncan, “Factors of safety and reliability in geotechnical engineering,”


Journal of geotechnical and geoenvironmental engineering, vol. 126, no. 4, pp.
307–316, 2000.

77
[45] K.-K. Phoon, Reliability-based design in geotechnical engineering: computations
and applications. CRC Press, 2008.

78
Curriculum Vitae

Senthilnath G T is born on 24th April 1986 in Tamilnadu, India. He started Civil


Engineering study at Anna University, India in Jul. 2003 and graduated in May 2007.
He pursued master’s degree in Soil Mechanics at Indian Institute of Technology, Roorkee
that involved DAAD (Deutscher Akademischer Austausch Dienst) fellowship at TU
Dresden, Germany and he graduated from it in Jul. 2009.

In Aug. 2009, he started working as Geotechnical design engineer and was mostly
involved in major infrastructure projects including Road and Rail Tunnels, Mining and
urban development projects.

In Jan. 2014, he started to pursue the 2nd level specializing master’s course at Politecnico
di Torino and after finishing the course work, he joined the consulting firm - Geoconsult
Asia Singapore as Executive Civil Engineer (Tunnels).

Contact: gt.senthilnath@geoconsult.com.sg, gt.senthilnath@gmail.com

79
Appendix A
List of stability models studied for this report.

Sl. No. Name of the Model Year Soil Type Analysis Type Type of Model Constitutive Model
3D model with linear triangular
1 Horn Model 1961 Homogeneous Global Equilibrium -
wedge

2 Murayama Model 1966 Homogeneous Global Equilibrium 2D log-spiral wedge Mohr-Coulomb material

3 Broms & Bennermark 1967 Homogeneous Emperical Model 2D/3D Empirical model Tresca material

2D/3D Lower & upper bound


4 Atkinson & Potts 1977 Homogeneous Stress Method Cohesionless material
plasticity solution - Cavity model
2D/3D Lower & upper bound
5 Davis method 1980 Homogeneous Stress Method Tresca material
plasticity solution

6 Kimura & Mair 1981 Homogeneous Emperical Model 2D/3D Empirical model Tresca material

2D/3D Lower bound plasticity


7 Muhlhaus 1985 Homogeneous Stress Method Mohr-Coulomb material
solution - Spherical opening
2D analytical model for stability
8 Ellstein 1986 Homogeneous Global Equilibrium Tresca material
number N
2D Quarter circle, Semi-circular &
9 Krause 1987 Homogeneous Global Equilibrium Mohr-Coulomb material
spherical model
3D Variational analysis model with
10 Mohkam Method 1989 Homogeneous Global Equilibrium Mohr-Coulomb material
log-spiral wedge
3D Lower & upper bound plasticity
11 Leca & Dormieux 1990 Homogeneous Stress Method Mohr-Coulomb material
solution - Conical Bodies
2D Empirical solution for Fracture
12 Mori 1991 Homogeneous Emperical Model Tresca material
/Blowout - NC soil

13 Chambon & Corte 1994 Homogeneous Stress Method 2D Empirical model Cohesionless material

14 Jancsecz & Steiner 1994 Homogeneous Global Equilibrium 3D Wedge model - Earth pressure Mohr-Coulomb material

2D Empirical solution for Fracture


15 Bezuijen 1996 Homogeneous Emperical Model Tresca material
/Blowout - OC soil
3D model with Linear triangular
16 Anagnostou & Kovari 1996 Homogeneous Global Equilibrium Mohr-Coulomb material
wedge with infiltration

17 Belter / Katzenbach 1999 Heterogeneous Global Equilibrium 2D linear wedge model Mohr-Coulomb material

3D Lower & upper bound plasticity


18 Lee & Nam 2001 Homogeneous Stress Method Mohr-Coulomb material
solution - Conical Bodies
3D model with linear triangular
19 Broere 2001 Heterogeneous Global Equilibrium Mohr-Coulomb material
wedge

20 Vermeer' Empirical Method 2001 Homogeneous Semi-emperical Model 2D/3D empirical plasticity solution Mohr-Coulomb material

2D plasticity solution obtained


21 Augarde et al 2003 Homogeneous Stress Method Tresca material
through FEM
3D Lower & upper bound plasticity
22 Caquot / C.Carranza-Torres 2004 Homogeneous Stress Method Mohr-Coulomb material
solution
3D Spherical cavity. Suitable for weak
23 Kolymbas Method 2005 Homogeneous Stress Method Mohr-Coulomb material
rock only
3D model with linear triangular
24 Girmscheid Method 2005 Homogeneous Global Equilibrium Mohr-Coulomb material
wedge

25 Klar et al 2007 Homogeneous Stress Method 2D/3D kinematic approach Tresca material

26 Mollon 2010 Homogeneous Stress Method 3D multi-block failure mechanism Mohr-Coulomb material

3D model with linear triangular


27 Anagnostou 2012 Heterogeneous Global Equilibrium Mohr-Coulomb material
wedge
3D upper bound plasticity solution -
28 Tang, Liu, Albers 2014 Heterogeneous Stress Method Mohr-Coulomb material
Layered soil
3D model with linear triangular
29 Perazzelli, Anagnostou 2014 Heterogeneous Global Equilibrium Mohr-Coulomb material
wedge with seepage forces
3D model with linear triangular
30 Chen, Tang, Yinm et al 2014 Heterogeneous Global Equilibrium Cohesionless material
wedge

80

View publication stats

Das könnte Ihnen auch gefallen