Sie sind auf Seite 1von 19

Engineering Geology 181 (2014) 93111

Contents lists available at ScienceDirect

Engineering Geology
journal homepage: www.elsevier.com/locate/enggeo

3D simulation of TBM excavation in brittle rock associated with fault


zones: The Brenner Exploratory Tunnel case
Kai Zhao a,, Michele Janutolo b, Giovanni Barla b, Guoxing Chen a
a
b

College of Transportation Science & Engineering, Nanjing University of Technology 210009, Nanjing, China
Department of Structural and Geotechnical Engineering, Politecnico di Torino, Italy

a r t i c l e

i n f o

Article history:
Received 18 March 2013
Received in revised form 27 June 2014
Accepted 3 July 2014
Available online 12 July 2014
Keywords:
Brenner Base Tunnel
Granitic rocks
Fault zone
Brittle failure
3D simulator
Thin rock pillar

a b s t r a c t
Brittle fault zones represent a major challenging geological environment for TBM tunnelling in deep Alpine
tunnels, particularly when the faults are near parallel to or cross the tunnel axis at a low angle. This is the case
of the Brenner Exploratory Tunnel in Italy. Serious local instabilities occurred at the left side wall during TBM
drive in the granitic rocks associated with a sub-vertical fault zone, parallel to the tunnel axis. The segmental
lining was collapsed at a distance of more than 2D (D is tunnel diameter) behind the face, without any evidence.
The deformation and failure then propagated intensively to nearby, previously stabilized sections with a length of
approximately 60 m in the longitudinal direction, leading to a subsequent damage of the shields and grippers of
the machine and to a stoppage of the excavation in almost 4 months.
To deal with these severe geotechnical problems encountered when tunnelling through a fault zone, a realistic 3D
numerical simulation based on a site investigation and characterisation of the fault zone, can provide a helpful
decision aid as they give a quantitative assessment of the potential mode of failure. In the case of the Brenner
Exploratory Tunnel, the behaviour of the rock mass is neither ductile nor brittle, but governed by the combination
due to the presence of the brittle fault zone. This paper focuses on the 3D simulation of such complex failure
evolution. Special emphasis is placed on the modelling of the fault zone and of the TBM excavation process.
The results demonstrate the role that the local rock mass condition and the complex interaction between the
rock mass, the TBM components, and the tunnel support play on the characterization of this instability
phenomenon.
2014 Elsevier B.V. All rights reserved.

1. Introduction
Brittle fault zones have been encountered in a great variety of underground projects. Extended reports exist in the literature, inter alia from
several tunnel cases in Turkey (Dalgic, 2000, 2003; Aydin et al., 2004),
the YacambuQuibor tunnel in Venezuela (Hoek and Guevara, 2009),
the Gotthard base tunnel and the Lotschberg base tunnel beneath the

Abbreviations: apeak, peak curvature exponent (HoekBrown); ares, residual curvature


exponent (HoekBrown); c, cohesion (MohrCoulomb); CI, crack initiation threshold; E,
Young's modulus; Kn, elastic normal stiffness of the interface element; Ks, elastic shear
stiffness of the interface element; Kt, elastic shear stiffness of the interface element;
mpeak, peak HoekBrown slope constant; mres, residual HoekBrown slope constant;
RMR, rock mass rating; speak, peak intercept constant (HoekBrown); sres, residual intercept constant (HoekBrown); t, time; T, tensile strength; u, displacement; uf, displacement
at the face; un, nal displacement; UCS, uniaxial compressive strength of intact rock;
UCSrm, uniaxial compressive strength of rock mass; p, hardening/softening parameter;
pi, principal plastic strains; pl, equivalent plastic strain; , friction coefcient; ,
Poisson's ratio; , stress; 1, maximum principal stress; 3, minimum principal stress; ,
shear stress; , friction angle (MohrCoulomb); , dilation angle (MohrCoulomb).
Corresponding author at: No.200, Zhongshan North Road, Nanjing, Jiangsu, 210009,
China. Tel./fax: +86 2558139212.
E-mail address: zhaokaiseu@yahoo.cn (K. Zhao).

http://dx.doi.org/10.1016/j.enggeo.2014.07.002
0013-7952/ 2014 Elsevier B.V. All rights reserved.

Swiss Alps (Kovari and Fechtig, 2000; Loew et al., 2010). However, it
still represents a major challenging geological environment for TBM
tunnelling in deep Alpine tunnels, particularly when the faults are
near parallel to or cross the tunnel axis at a low angle. This is the case
of the Brenner Exploratory Tunnel in Italy. Serious local instabilities occurred at the side wall during TBM drive in the granitic rocks associated
with a sub-vertical fault zone parallel to the tunnel axis, leading to a signicant damage of the shields and grippers of the machine and to a
stoppage of the excavation in almost 4 months. The present paper
gives a brief historical review of the construction of this exploratory tunnel and, particularly of a geomechanical characterization of the tunnel,
with main concern of the situation around chainage 6 + 151.
At the principal issues to overcome this adverse geological condition
is to understand and correctly anticipate this abrupt failure. During TBM
advancements, except for the primary deformation induced by the
excavation, some unexpected deformation was observed at areas
remote from the face, and lead to large-scale failures. Based on the
monitoring convergence data, Kontogianni et al. (2004), Kontogianni
and Stiros (2005, 2006), and Kontogianni et al. (2008) revealed that
this additional deformation has a systematic, clear pattern and shares
some common characteristics. Yet, difculties are met in making
reliable predictions at the design stage. In the present paper, we try to

94

K. Zhao et al. / Engineering Geology 181 (2014) 93111

investigate this deformation mechanism and provide a quantitative


assessment tool, by means of completely 3D numerical modelling.
A 3D nite element model for TBM tunnelling in squeezing and
spalling rock conditions, which takes into account all relevant components and realistically models the step-by-step construction process, is
proposed in Zhao et al. (2012). The present paper extends this 3D
model, with potential to account for the discontinuous behaviour of
the rock mass (with fault zones) and also for the relevant interaction
process between the ground, the TBM and the support systems, with
reference to the Brenner Exploratory Tunnel case.
2. Project description
2.1. The Brenner Base Tunnel
The Brenner Base Tunnel is the main infrastructure to be built along
the railway corridor between Munich (Germany) and Verona (Italy),
along the NorthSouth axis of the European high-speed (for passengers)/high-capacity (for freights) rail link HelsinkiValletta (European
Commission, 2011). The construction of the Brenner Base Tunnel will
encourage the development of intermodal transport in the Alps, by
shifting freight and passengers to the railway.
With a circular internal cross-section of 4.05 m radius, the Brenner
Base Tunnel will be 55 km long. It is a twin-tube system consisting of
two single track tunnels, connected together by cross passages at 333 m
intervals and built 40 to 70 m to each other according to the geological conditions of the rock mass. TBM excavation is likely to be adopted
for 70% of the total length. A specic feature of this tunnel is the planned
construction of a pilot tunnel, generally 10 m below the two main tunnels to be used later for drainage and emergency access. A 6.3 m diameter exploratory tunnel excavated on the Italian side between Aicha and
Mules has been completed in 2010. In Austria, a pilot tunnel between
Innsbruck and Ahrental, started in 2009, has also been completed
(Figure 1).
2.2. The Aicha exploratory tunnel
2.2.1. General features
The Aicha exploratory tunnel, as shown in Fig. 2, has been excavated
from April 2008 to November 2010 by a 6.3 m diameter Double Shield
Universal (DSU) TBM. This exploratory tunnel does not follow in
the rst part the main tunnel alignment, as the portal of Aicha is
located further south with respect to the main tunnel portal. Until
chainage 7 + 835 it is aligned with the main tunnel, lying in the middle
of the two tubes 10 m below.
Simultaneously with the excavation, a 20 cm thick segmental lining
has been installed. It was assembled by ve precast concrete segments
with L = 1.5 m, D = 6 m (L = length, D = diameter), plus the closure
oor segment containing a canal for water drainage. As soon as the
segment ring was positioned, the void between the lining and the
rock mass was lled by pea-gravel (without mortar injection).
In order to update the knowledge of the rock mass conditions and
to monitor ground-TBM interaction during excavation, a systematic
investigation programme was necessary for the TBM, with the following
equipment:
Several windows for inspection/entry at the face, for the observation
and classication of the encountered rock mass
Windows in the shield for the direct systematic observation of the
walls
Probe drilling ahead of the face
A measure and registration system of the excavation operative
parameters.
Furthermore, geo-seismic investigations and geo-electrical surveys
were conducted continuously, as well as other monitoring activities
(Grandori et al., 2011).

Fig. 1. New stretches along the existing railway corridor MunichVerona (from www.
bbtinfo.eu).

The geology in the project area is characterised by Brixner Granite


tectonic unit of the Southern Alps (Figure 3). It is a non-metamorphic
unit which underwent the main deformation during the Alpine orogeny. The most commonly encountered rocks are biotitic granites,
biotiticamphibolitic granodiorites and local leucocratic varieties.
Different sets of joints and faults (mainly subvertical) cross-cut this
homogeneous lithology, which are mainly responsible for the different
rock mass qualities encountered during the tunnel excavation, associated with water inows. Brittle deformations including these joints and
faults occurred along the tunnel alignment. Signicant deformation
was observed in the fault zones, determined by the unfavourable orientation and position. In particular, the most serious instability events occurred in the sector around chainage 6 + 151, where the fault zones are
parallel and tangent (almost touching) to the left tunnel wall along the
tunnel axis (discussed in detail in Section 3).
The behaviour of the rock mass was, in general, stable outside the
fault zones. Only slight gravity-driven instability (i.e., spalling and
rockburst) were encountered along the stretch (Grandori et al., 2011).

K. Zhao et al. / Engineering Geology 181 (2014) 93111

95

Fig. 2. Aicha exploratory tunnel and Mauls adit (from www.bbt-se.com).

2.2.2. Construction history


After the rst 2 km in very hard, massive and abrasive rock mass,
the exploratory tunnel entered the stretch (with a length of 3.5 km),
characterised by alternating zones with intact rocks and zones with
fractured rocks (with minor length) (Grandori et al., 2011). In the rst
5.5 km stretch, the uniaxial compressive strength (UCS) of the intact
rock was around 154 MPa and the rock mass was mainly categorized
as class RMR II, except in some fault/fractured zones (with limited effects). In the following section, between chainage 5 + 500 and chainage
6 + 200, the granite exhibits highly fractured and altered. An example
of the detailed RMR rock mass classication, on the basis of measurements and observations conducted on the representative rock mass
condition of this fractured granite, is given in Table 1. In particular, the
faults causing serious instability phenomenon are located in this section, which are the main concern in this paper. Since then, the exploratory tunnel arrived at the zone of massive or slightly fractured rock
mass, extending to the end of this tunnel. The UCS was lower than
that in the rst part, around 120 MPa, probably as the Periadriatic Line
was approaching.

a sudden increment of the deformations on the day the failure occurred


at chainage 6 + 151. It is an entirely extraordinary event with respect to
that observed in the previously installed monitoring station, and obviously, not as a result of accumulating deformation due to the face effect. Total strain at the monitoring sites reached 900 in the
extrados (i.e., zone of compression) and 600 in the intrados (i.e.,
zone of tension) at the left-hand sidewall lining (9 'o clock position)
(Figure 4). The corresponding stress was 35 MPa in compression and
20 MPa in tension, inducing permanent damage of the lining in this section. Since then, it is seen that the stress redistribution occurs within the
damage area. The maximum stresses decrease to the values of 30 MPa
and 15 MPa, respectively, which indicate a tendency of crack propagation along the longitudinal direction (Figure 5).
The results of systematic measurements at this section for the 3 daylong subsequent period (August 7th to August 10th) indicated a tendency for stabilization. With TBM advancement and the consequent
increasing distance from the face, the deformation continued to accumulate till August 10th, 2009 when the excavation was arrested.
The following conditions were encountered after the TBM stopped
(Barla et al., 2010):

3. Stability problem

The rock mass at the face was scarcely fractured and composed of a
non-altered granite; the only exception was a very fractured zone
localised at the bottom left.
The shield was in contact with the rock mass along the left side
almost up to the face, while it was completely detached from the
bored prole on the right side.
The gap between the lining rings and the bored prole was correctly
lled with pea gravel on the right side, while it was almost completely
absorbed by the deformations on the left side for the entire stretch
affected by the phenomenon.
A few segments of the left-hand lining sidewall were wrecked so that
the rock mass was visible at the back: cloritized cataclastic granite,
with angular decimetric clasts (Figure 6a).
Breaking and cracking of the lining were mainly sub-horizontal and
were concentrated in the portion of the lining around the horizontal
diametral plane (Figure 6b). Cracks affected directly the segments.
The right-hand sidewall experienced neither damage nor important
displacements; a windowed segment showed a hard and unaltered

3.1. Instability phenomenon at chainage 6 + 151


The instability phenomenon at chainage 6 + 151 started on August
7th, 2009 when damage was observed at the left-hand sidewall lining
during TBM advancement. In the following days the damage extended
towards the face, reaching the shield, and also backwards, progressively
decreasing to a fracturing state up to approximately 60 m behind the
face. After 710 days, this sudden phenomenon stabilized.
Prior to the failure, as well as during and after the rehabilitation
works (Section 3.3), the tunnel was systematically monitored on the
basis of the extensometers. The instruments were installed with vibrating wire in the reinforcing bars at the extrados and intrados of the segments at certain positions. In particular, among them, the section at
chainage 6 + 099 was installed on August 5th, 2009 (two days before
the failure occurred), which was 52 m behind the breakthrough point
of the tunnel (chainage 6 + 151). Recordings at this section revealed

K. Zhao et al. / Engineering Geology 181 (2014) 93111

Fig. 3. Geological section along tunnel alignment.

96

K. Zhao et al. / Engineering Geology 181 (2014) 93111

97

Table 1
RMR classication parameters and their rating parameters for the representative fractured granite.
Classication parameters and their ratings
Parameter

Range of values
N10 MPa
N250 MPa
15
90100%
20
N2 m
20
Discontinuity length (Persistence) b1 m
Rating
6
Separation (Aperture)
None
Rating
6
Roughness
Very rough
Rating
6
None
Inlling (gouge)

410 MPa
100250 MPa
12
7590%
17
0.62 m
15
13 m
4
b0.1 mm
5
Rough
5
Hard lling b 5 mm

24 MPa
50100 MPa
7
5075%
13
200600 mm
10
310 m
2
0.11.0 mm
4
Slightly rough
3
Hard lling N 5 mm

12 MPa
2550 MPa
4
2550%
8
60200 mm
8
1020 m
1
15 mm
1
Smooth
1
Soft lling b 5 mm

Rating
Weathering
Rating
General conditions
Inow per 10 m tunnel length
Rating

6
Unweathered
6
Completely dry
None
15
Very favourable

4
Slightly weathered
5
Damp
b10 l/min
10
Favourable

2
Moderately weathered
3
Wet
1025 l/min
7
Fair

2
Highly weathered
1
Dripping
25125 l/min
4
Unfavourable

10

Strength of intact rock material Point load test


UCS
RQD
Spacing of discontinuities
Condition of discontinuities

Groundwater

Discontinuity orientations

RMR index: 45
Rock mass class: III

Fig. 4. Monitored strain in the lining at chainage 6 + 099.

525 15 b1
2
1
0
b25%
3
b60 mm
5
N20 m
0
N5 mm
0
Slickensided
0
Soft
lling N 5 mm
0
Decomposed
0
Flowing
N125 l/min
0
Very
unfavourable
12

98

K. Zhao et al. / Engineering Geology 181 (2014) 93111

Fig. 5. Stress in the lining at chainage 6 + 099.

granite.
The collapse of the lining on the left side blocked the segment carrier
and deformed the back-up, drifting the shield with respect to the
lining (Figure 7).
Furthermore, it is interesting to note that another failure occurred at
chainage 6 + 105 and chainage 6 + 112 on August 10th (i.e., after the
stabilization of the serious failure events at chainage 6 + 151). It was
caused by the execution of a probe borehole at the left sidewall. The
fracture extension and propagation induced serious damage of 4
segments. The borehole was intercepted by a pressure water inow at
20 m from the left sidewall, i.e., at 13 m from the fault zone. It reected
that no water was contained in the fault zone, which was also veried

by the observation at chainage 6 + 123 that the breccia at the end of


the lining was dry.

3.2. Geological model


The original site investigations involved walk-over surveys and core
drilling. In total, 12 boreholes were drilled and allowed one to reconstruct the geological model, as shown in the plan-view of Fig. 8
and in a cross-section of Fig. 9 at the tunnel level. The boreholes denoted
the presence of a fault dipping at high angle towards NE (N 5060),
running parallel and tangent (almost touching) to the left tunnel wall
between chainage 6 + 020 and chainage 6 + 151.

Fig. 6. Lining breakage at chainage 6 + 151: (a) Disjointed segment immediately behind the shield with outcrop of cloritized cataclastic granite (Barla et al., 2010); (b) Sub-horizontal
cracks in the lining.

K. Zhao et al. / Engineering Geology 181 (2014) 93111

99

Fig. 7. Effects of the deformations of the segments on the back-up: (a) carpentry injured; (b) segment detached from the shield; (c) segments' belonging.

The overall thickness of the fault zone ranged from 7 to 10 m. It


consisted of a core of non-cemented cataclastic rocks with an average
thickness of 5 m and of an external damage zone with highly fractured
granite (Figure 10). The fault was crossed/dislocated by minor bundles
of faults.
In addition to this core, other minor faults were detected at 1430 m
to the left sidewall. In the other stretches the boreholes exhibit few to
meanly fractured, with an alteration along the discontinuity surfaces.
According to the observations at proximity of the cracked lining and in
the injection holes for the pea gravel, the fault breccia could originally
be very compact and dry with a thin silt component. This fault material
had a residual cohesion before the deconnement due to the excavation, as a result of compaction for the charge and crystallisation of
mineral sin-kinematic phases during the fault movement (chlorite,
quartz, epidote). Clay gouges were detected only in some short boreholes with a maximum thickness of 5 cm.
From the geological model, it is seen that the size and orientation of another fault zoneRio Bianco dominates the morphology,
as it determined the deep incision of the homonymous stream. It is
composed of a core zone and a damage zone, as it can be observed
in surface (Figure 11). Nevertheless, this fault has been crossed
by a more favourable angle and lower topographical cover. For
this reason, it did not cause signicant stability problems during

the excavation as observed, except the heavy water inows


encountered.
Furthermore, the geo-structural kinematic analysis has been performed in the outcrops on the eld. The fault was interpreted as a
particular element strictly related to the Rio Bianco fault zone: as
schematically illustrated in Fig. 12, it is a lateral strikeslip fault conjugate to the Rio Bianco fault. The two faults are related by the structural elements of the Brixner Granite which developed in the frame
of a stress eld, with a NWSE principal compression axis. Therefore,
the fault was considered as a local disturbance and it was expected
that similar features could not be met in the other parts of the rock
mass, as actually observed. After chainage 6 + 151, the fault zone
gradually diverged from the tunnel axis and then disappeared,
going towards North in the direction of excavation.
This geological model was in perfect agreement with the observations when the excavation re-started.
3.3. Rehabilitation works
First of all, several activities (i.e., bolting, installation of steel ribs,
etc.) in order to ensure the safety of the tunnel were undertaken.
Then, the following works were carried out to restore the damaged
tunnel:

Fig. 8. Geological reconstruction (plan-view) (Barla et al., 2010).

100

K. Zhao et al. / Engineering Geology 181 (2014) 93111

permanent deformation on the entire left side with considerable


denting on the lower zone
Filling of the tunnel with pea gravel.
3.4. Resume of the excavation
On December 4, 2009, reasonable progress was made after introducing a special segmental ring. It was formed with the same segments of
the previously installed rings, but with additional reinforcement and
steel crown segments.
For the continuation of the advancement outside the fault zone,
several options for the support design were suggested to increase the
order of load capacity, listed as follows (Barla et al., 2010):

Fig. 9. Reconstruction of the geological section at chainage 6 + 100 (Barla et al., 2010).

a) Installation of standard concrete rings reinforced with 10 mmdiameter steel bars


b) Installation of concrete rings reinforced with UPN 120 structural
steels
c) Installation of concrete rings reinforced with structural steels and
with guide bars, in order to sustain heavily asymmetric loads
d) Installation of concrete rings with guide bars and reinforced with
structural steels also between two adjacent rings.
3.5. Risk management for the remaining stretch

Injections of lling and stabilizing resins


Cutting, demolition and replacement of reinforced concrete segments
showing excessive shifts and/or breakage, for subsequent smaller
portions, with steel panels, subsequent rear lling of panels with
cement and installation of radial nails with 3 m long self drilling bars
Dismounting of the portion of the segment carrier which was not
blocked by the shifts in the lining
Excavation of an unlocking tunnel for the TBM (Figure 13)
Restoration of the gripper: the grippers showed a signicant deformation of both shoes, a puncture of the front part of the shield and the
collapse of a supporting ank
Restoration of the damaged shield: the shield showed widespread

After the worst hazard scenario occurred in chainage 6 + 151, it is


recognized that the problems encountered in tunnelling through
the brittle fault zone are so diverse and complex that monitoring can
be employed meaningfully, as well as their subsequent evaluation.
Therefore, the Geotechnical Risk Management Procedure was suggested
within the framework of the observational method. Monitoring was
applied in different ways. In particular, the load history on the lining
was continuously recorded during TBM advancement, and also the
following data were considered to improve the decision making on
selecting appropriate support system and to establish early warning
systems against incipient collapse:

Fig. 10. Boring at chainage 6 + 117 (05 m) and 6 + 060 (05 m): tectonic breccias.

K. Zhao et al. / Engineering Geology 181 (2014) 93111

101

Fig. 11. Outcrop of the fault in the Rio Bianco Valley.

Rock mass classication (weighted RMR on the excavation face)


Important asymmetries in the rock mass structure at the face
Data of the machine (Grandori et al., 2011)
Analysis of the muck
Water inow measures.

In addition, the monitoring of the machinerock mass interaction


was performed based on the analysis of the TBM operation data. In a
good to excellent rock mass condition, the thrust value and the penetration rate were around 10 MN and 20 mm/min, respectively. The strong
variations of these monitoring data could denote a worsening of the
rock mass quality, which imposed a further face mapping in order to decide either to maintain or to change the support type (Barla et al., 2010).
In reality, as the excavation was carried out by a Double Shield TBM, the
tunnel faces could not be documented continuously using rock mass
classication systems, but only with a frequency.

Furthermore, the following parameters were systematically introduced for the Geotechnical Risk Evaluation (Barla et al., 2010):
the re-gripping pressure: an anomalous value can denote an anomalous load on the tail shield and therefore on the lining
Volume of pea gravel injected instead of the theoretical value: it can
denote the rock mass deformations
Values of the induced deformations/stresses in the lining. It is important to dene the threshold values for the strain in the lining
(attention value, alarm value).
With the aid of these measurements, one is in the position to keep
the ground pressure under control and the lining resistance can be
chosen accordingly, until an optimum value is achieved.
4. Computational model
4.1. Foreword
In this paper, a simulator of TBM excavation of deep tunnels has
been developed to this end, by using 3D FEM modelling and the midas
GTS (Geotechnical and Tunnel Analysis System) computer code (TNO
DIANABV) (Midas, 2010), to reproduce and to analyse the instability
phenomenon as described before. The presence of water pressure and
consolidation problems is not taken into account. Emphasis is placed
on the discontinuous behaviour of the rock mass and on the interaction
between the rock mass and the TBM and the support components
associated with the progressive advance of the working face.
4.2. Model construction

Fig. 12. Structural scheme which describes the relationship between the Rio Bianco fault
and the causive fault encountered in the tunnel; the bigger arrows show the palaeo-axis
of tectonic deformation which determined the movements (little arrows) along the fault
network of the zone (Barla et al., 2010).

The geological conditions have been simplied by considering only 2


rock types: the fault and the granitic rock mass. The fault is assumed to
be parallel to the excavation, 7 m thick and approximately 0.2 m far
from the excavation boundary. The dip of the fault has been assumed
to be 73 N towards NE. Due to the highly heterogeneous behaviour of
the rock mass, the entire domain is modelled. A cylindrical domain
is used in order to acquire high mesh quality. 8-node hexahedron
solid elements are chosen as appropriate. In particular, a high level of
mesh renement is desirable to capture high strain gradients of the
pillar rock between the side wall and the fault zone. The transition
elements are used in order to assure a better shape at the interface between the fault zone and the host rock mass. Fig. 14 shows the mesh layout of the model.
The mesh size should be large enough so that the external boundary
can represent an innitely extended medium. Otherwise, the presence
of the articial boundaries will induce a signicant inuence on the
stressstrain eld around the tunnel. The external boundary is set to a

102

K. Zhao et al. / Engineering Geology 181 (2014) 93111

Fig. 13. Unlocking tunnel for the TBM: (a) photo; (b) cross-section (Grandori et al., 2011).

distance of 11D in the transversal plane to minimize boundary effects. In


the longitudinal direction, a total length of 20D is applied in order to
maintain a sufcient distance between the rear boundary and the last
excavation face. The mesh discretization is equal to the excavation
length (1 m); then, after the last excavation face, it is gradually
increased. In all cases, parametric studies are necessary to evaluate the
appropriate size for every specic case.
As far as mesh grading in the transversal plane, the accuracy of the
results depends on the ability of the mesh layout to capture the high

strain gradients in the overstressed zone of the rock mass (Diederichs,


2007). In this failure localization cases, a very rened region around
the tunnel contour with an element size of 20 20 cm at the tunnel
boundary on the left side and a coarser zone away from it are used.
Once the 3D model is created, the following boundary conditions are
imposed:
In the circular outer boundary, displacements along the vertical
(Y) and the horizontal direction (X) are prevented;

Rock mass
Fault

Fig. 14. 3D rock mass mesh (initial stage).

K. Zhao et al. / Engineering Geology 181 (2014) 93111

103

In the front and in the rear outer faces, displacements along the
excavation advance direction (Z) are prevented.
As the model refers to a deep tunnel with a signicant fault close to
the boundary, the in situ state of stress is applied as a hydrostatic initial
stress without consideration of the free ground surface and of the stress
gradient due to the gravity. In reality, in the Brenner tunnel case, the
discontinuities, i.e., fault zones, act to signicantly perturb the stress
eld (i.e. local effect) and thus the eld measurements become an
essential component of the overall design process. The in situ stress
measurements in the boreholes showed that the range of k0 (ratio of
horizontal stress/vertical stress) is comprised from 0.8 to 1.2. Also by
taking into account the presence of the pre-sheared fault zone (Hoek
and Marinos, 2010), it is assumed that the horizontal and vertical
stresses are equal in the 3D model. The in situ state of stress is thereby,
assumed to be with a vertical and horizontal state of stress of 12.72 MPa,
evaluated at 480 m depth, by introducing a rock mass unit weight equal
to 26.5 kN/m3.
4.3. Rock mass model
The stability problems with respect to the discontinuous nature of
ground are not only a result of brittle fracture of the hard rock at great
depth, but also of shear failure along pre-existing discontinuities. The location and properties of the fault zone are of paramount importance and
therefore, respected in the 3D model in particular to identify the
potential failure mode. Furthermore, two constitutive models for reproducing the discontinuous behaviour of the rock mass in a consistent
and simple way are illustrated in the following.
4.3.1. Granitic rock mass (brittle failure)
Brittle failure, both in the form of major spalling and potentially
strain burst, often dominates rock damage and failure processes in crystalline rocks (like granitic rock mass) near excavation boundaries under
high in situ stress environment (Diederichs, 2003). In the case of the

Fig. 16. Granitic rock mass, Diederichs model (based on the generalised HoekBrown criterion) and equivalent one (based on the traditional HoekBrown criterion).

Aica exploratory tunnel, the low connement caused by both the presence of the excavation boundary on one side and the fault a noncohesive material on the other side, generates a brittle failure typical
of hard rocks resulting in progressive slabbing and spalling processes.
Due to the fact that brittle failure involves a tensile fracturing
process, the conventional yielding criteria for continuum medium as
the MohrCoulomb and the HoekBrown cannot properly describe
the actual behaviour of the rock (Kaiser et al., 2000). In this respect,
Diederichs (2007) proposed a specic criterion for susceptibility to brittle spalling (as opposed to plastic shear), based on the generalised
HoekBrown criterion and an elasticperfectly-brittleplastic constitutive

Fig. 15. Composite strength envelope for brittle rocks (Diederichs, 2007).

104

K. Zhao et al. / Engineering Geology 181 (2014) 93111

Table 2
Model parameters for Granitic rock mass, fault zone and the interface between the fault
zone and the granitic rock mass.
Granite
General parameters

Criterion of
Diederichs

Equivalent parameters

UCS
T
CI

Fault zone
E

UCS
GSI
mi

25 GPa
0.25
140 MPa
8.1 MPa
63 MPa

apeak
speak
mpeak
ares
mres
sres

0.25
0.041
0.656
0.75
9
106

250 MPa
0.25
7 MPa
25 5
19

0
38
0

apeak
mpeak
sres
ares
mres
sres
Interface
Kn
Ks
c

0.5
1.4
0.2
0.5
6
106
10 GPa/m
1 GPa/m
0
38
0

law. Peak and residual yield functions are dened by damage


threshold and spalling limit, respectively (Figure 15).
A model based on the traditional HoekBrown criterion equivalent
to the generalised one proposed by Diederichs (2007) is adopted, as
shown in Fig. 16. The procedure for determining the input parameters
for the generalised HoekBrown criterion is the following:
Determine the crack initiation threshold CI from uniaxial compression
tests;
Set apeak to 0.25;
Obtain a reliable estimate of tensile strength, T (from laboratory
tests);
Calculate the appropriate s and m from:

speak CI=UCSapeak

mpeak speak UCS=jT j

Set ares = 0.75, sres = 0 or 106 (for numerical stability) and mres = 5
to 9 in order to model the transition envelope to high connement
shear (spalling limit).
In particular, CI is empirically estimated as 0.45 UCS according to Cai
et al. (2004), who suggest that the ratio between CI and UCS is generally
within the range of 0.30.5. The other parameters are obtained according to Diederichs (2007).
The assumed rock mass parameters are shown in Table 2. It
should be noted that the model cannot consider shear at high connement correctly and thus it is only limited to the near-excavation
analysis.
4.3.2. Fault zone
Alpine fault zones are complex structures, exhibiting highly heterogeneous rock mass conditions. The simulation of the behaviour of the
fault zone must take into account both (i) shear across a cataclastic
crush zone in the fault zone, and (ii) sliding on existing planes between
the fault zones and the granitic rock mass.
For the shear behaviour of the fault zone, we adopted the standard
linearly elastic, perfectly plastic material model with the MohrCoulomb
yield criterion and non-associated ow rule. The weak cataclastic rock
masses in the Brenner Exploratory Tunnel fail in shear through the mass
rather than along individual discontinuities, which can be traced back to
the fault zones down to depths at which the frictional resistance to slip

exceeds the ductile yield strength of the rock mass. However, it should
be kept in mind that fault zones represent complex geological structures, which are composed of various rocks with very different material
properties. The geomechanical properties of the faulted rock mass is
hard to characterise because of many difculties arising in getting
representative samples during eld investigations, in specimen preparation, and in performing appropriate laboratory testing. Therefore,
in addition to the laboratory testing on the core samples, the key parameters were empirically estimated in the eld on the basis of the descriptive classication (GSI) of Hoek et al. (1998). In particular, the GSI chart
for heterogeneous rock masses is used to account for the lling materials in the fault zones. Additionally, a weighted average of the intact
strength properties of the strong and weak materials is considered as
the reasonable approximation of the equivalent material properties
UCS and mi for the fault zones, as proposed for highly heterogeneous
rock mass by Marinos et al. (2006). Based on the estimated UCS, the
material constant (mi) value and GSI value attributed to this faulted
rock mass, the cohesive strength and friction angle were also evaluated
for design and analysis purposes (Barla et al., 2010). The parameters are
shown in Table 2.
On the other hand, zero-thickness interface elements are used
and the criterion for Coulomb frictional law is assumed to represent the potential sliding mechanism between the fault and the
rock mass. The elastic and plastic parameters (according to the
Coulomb friction law) are listed in Table 2. It is assumed that the
interface has the same friction coefcient as the fault material as
well as no cohesion.

4.4. Main TBM components


The following TBM components are considered:
Cutterhead, which is modelled with plate elements (with the stiffness
properties of steel) at the current excavation face and where a pressure is applied, given by the maximum cutterhead thrust divided by
the area;
Shields (front, telescopic, gripper and tail shield), which are modelled
with plate elements applied at the excavation boundary and with the
stiffness properties of steel;
Pea gravel, which is modelled with solid elements and is activated 2 m
behind the shield;
Grippers, which are modelled indirectly, as a portion of the shields in
which a given pressure is applied;
Lining, which is modelled with plate elements (with the stiffness
properties of concrete); the segmentation of the tunnel lining is
disregarded (the lining is assumed as continuous), since the model is
not especially designed to investigate in detail the behaviour of the
tunnel lining.
All these components have a linearly elastic isotropic law. The TBM
layout for the simulation is shown in Fig. 17. The size of the gap between
shield and ground, as well as between segmental lining and ground, is
taken as constant around the circumference. In reality, the shield slides
along the tunnel oor, which means that the gap around the shield and
segmental lining is not uniform and wider above the centre than in the
lower portion of the tunnel cross-section. Given the aim of this paper
(to simulate the instability which occurs mainly at the sidewall with
the gap of average size), however, it is sufcient to only consider the
contacts between the shield and sidewall rock mass.
The simulation parameters are given in Table 3. The elastic modulus
of the lining is the composite modulus of the reinforced concrete: the
concrete has class C45/55 (E = 36 GPa) and the reinforcement consists
in 8 mm-diameter bars with a 10 cm spacing. The self-weight is also
applied to all these components.

K. Zhao et al. / Engineering Geology 181 (2014) 93111

105

Fig. 17. TBM layout considered (plan-view).

4.5. TBM advancement and simulation procedure

4.6. Interaction between the machine components and the rock mass

A step-by-step method is adopted to simulate the TBM advancement.


The excavation length is taken to be equal to 1 m in order to:

The simulation of the interaction problems requires a special


attention to the discontinuous behaviour at the following interfaces:
(i) rock mass-shield and (ii) pea gravel-lining. This behaviour involves
frictional sliding and the possible closure of the gap between the shield
and the rock mass.
Zero thickness interface elements are used to represent the frictional
shearing mechanism. High normal and shear stiffnesses have been set.
For plastic slip, a Coulomb friction law with null-cohesion and a value
of the skin friction coefcient = 0.3 have been adopted for both the
interfaces. Since the numerical formulation used is based on the smallstrain/displacement assumption, the shields have been modelled right
on the tunnel boundary even if there is a gap in-between. In order to
simulate the gap between the rock mass and the shields, special interface elements have been used in which the elastic stiffnesses are set to
zero, i.e., Kn = Ks = Kt = 0, so that no stress can transfer from the
rock mass elements to the shield elements.
In order to know the regions along the shields in which the contact
with the rock mass takes place (on the left sidewall in this case), the
eld of the radial displacements has to be carefully monitored. When
these displacements exceed the radial gap, the properties of the interfaces in the relevant step are modied. As a slice of the rock core ahead

reproduce as well as possible the continuous process of TBM


excavation (Vlachopoulos and Diederichs, 2009);
improve the convergence of the elasto-plastic solution;
trade-off accuracy and computational time.
The construction stages are dened as follows:
in the rst step, initialisation takes place accounting for the in situ
stress eld;
in the second step, the TBM enters the model (the cutterhead
elements are activated) and the rst slice is excavated;
in the further steps, the rst slice of the front shield is activated,
according to the following paragraph;
in each step thereafter the TBM (front and rear shield) progresses into
the model;
in the fteenth step, the lining is activated as well as the pea gravel;
the stages proceed until a steady-state condition is reached.
Fig. 18 illustrates the simulation process on the left half of the
domain.

Table 3
Parameters for the components of TBM excavation.
Cutterhead

Shield

Diameter
Thickness

6.3 m
3 cm

Young's modulus
Poisson ratio
Pressure at the face

200 GPa
0.3
0.4 MPa

Lining

Grippers

Diameter
Thickness
Length front shield
Length rear shield
Young's modulus
Poisson ratio

6.3 m
3 cm
5m
7m
200 GPa
0.3

Outer diameter
Thickness

6m
20 cm

Young's modulus
Poisson ratio

38.1 GPa
0.2

Friction coefcient rock-shield skin

0.3

Friction coefcient rock-lining

0.3

Base
Height

Pressure

Pea gravel
2m
4.5 m

7 MPa

Thickness
Thickness

15 cm
15 cm

Young's modulus
Poisson ratio

1 GPa
0.3

106

K. Zhao et al. / Engineering Geology 181 (2014) 93111

(a) 13th step

(b) final step


Fig. 18. Illustration of the 3D model on the left side at two excavation steps. The two gradations of blue differentiate the parts of the shield in contact with the rock mass (dark-blue) with
the parts which are not in contact (light-blue).

of the face is excavated, the displacement uf at the face is removed.


To consider the geometry update, the nal horizontal displacement
un to be monitored is un = u uf, where u is the horizontal
displacement.
Furthermore, the gripper shield is activated as it is in contact with
the rock mass through the grippers: the properties of the interfaces
allowing the contact are therefore activated for the size of the gripper
shoes.

5. Results and discussions


Evidence from the Brenner Exploratory Tunnel indicates that, except for the primary deformation (which is a consequence of the spatial stress redistribution associated with progressive excavation),
under certain circumstances additional, unexpected deformation is
observed at areas far from the working face, and can even lead to
large-scale failures. This deformation and also the induced collapse

K. Zhao et al. / Engineering Geology 181 (2014) 93111

of tunnel segments are explained with the numerical results in this


section. It starts with a discussion of the local instabilities along the
tunnel (Section 5.1), and illustrates how the interaction between
the shield, the ground and the tunnel support inuence the degree
of overstressing of the ground (Section 5.2). Specically, this section
also shows the load transfer along the longitudinal direction, which
shed some light on this extensive deformation remote from the tunnel face (Section 5.3).

107

pillar between the tunnel wall and the fault. This discontinuous stress
distribution in the vicinity of the boundary between the weak and stiff
rock can be traced back to the strong stiffness contrasts between the
adjacent host rock and the fault zone. This stress concentration leads
to the brittle failure of the more competent pillar rock, associated with
the complex interaction between the ground, TBM and support system
at the excavation side (Section 5.2). Furthermore, a stress redistribution
can be checked in the fault zone, which is responsible for the yielding of
the weak fault material.

5.1. Local instabilities


5.2. Interaction between TBM, ground and tunnel support
First of all, the failure zones denoting the collapse of the thin rock
pillar between the tunnel wall and the fault are analysed. The degree
of failure is expressed through the equivalent plastic strain, dened as:
pl

r

2 2
2
2
p2 p3
3 p1

with p1 , p2 , and p3 the principal plastic strains.


This denition comes from the most commonly used expression for
the softening/hardening parameter (based on incremental plastic
strain) which is (Vermeer and De Borst, 1984):
p

r

2
p2 p2 p3 p3 :
3 p1 p1

It is noted that this softening parameter has not been introduced in


the constitutive law, and the equivalent plastic strain is thus only a
variable describing the degree of damage in the rock mass.
Fig. 19 shows such plastic strain contours in a 3D view of the left side,
while Fig. 20 in a cross-section at the steady-state. As in this section
dominated by the weak fault zone on the left sidewall, the failure is
localised, in the manner of the combination of the brittle failure of the
pillar rock between the tunnel and the fault, and the shear yielding of
the gouge. Furthermore, as the breakage of this pillar, the yielding
fault zone is loosened and thus the weak gouge material washed out.
This coincides very well with the in situ observations.
Fig. 21 depicts the maximum principal stress contours around the
excavation. It is seen that a signicant stress is channelling into this

Fig. 22 provides a complete picture of the horizontal displacements


on the left sidewall of the tunnel along the longitudinal direction. The
nal displacements, as already described in Section 4.4, are obtained
by removing the displacement at the face uf (the so-called predeformation). The nal displacements along the line corresponding to
the minimum thickness of the rock pillar between the tunnel and the
fault, are plotted in Fig. 23, as well as the ground pressure developed
on the shield and the lining. It is noticed that: (i) the displacements
close the gap at the front shield (equal to 3.5 cm) 2 m behind the face,
(ii) the contact zone of the rear shield through the grippers, together
with the applied pressure, blocks the displacement in these areas, (iii)
the displacements close the gap at the tail of the rear shield (equal to
9.5 cm), but for a really short distance (around 50 cm), so that the last
slice of the rear shield has not been activated (since the discretization
is 1 m), (iv) the displacements ceased when the support system is
installed. The maximum nal horizontal displacement at the steadystate on the left side is equal to 11.3 cm, while on the right side is only
equal to 0.5 mm.
In reality, not only the nal magnitude of the displacements is of
importance, but also the development with TBM advancement, as this
controls the shield and lining load development. Fig. 23b shows the
distribution of the ground pressure acting on the shield and the lining,
as soon as the converging ground closes the gap. Unexpectedly, a signicant load concentration develops at the end of the front shield. The
segmental lining also experiences a dramatic increase of the ground
pressure with the distance from the tunnel face even as the face effect
becomes less pronounced.

Fig. 19. Plastic strain contours along the tunnel in a 3D view on the left half of the model.

108

K. Zhao et al. / Engineering Geology 181 (2014) 93111

Fig. 20. Plastic strain contours in a cross-section at the steady-state.

It should be noted that the prefabricated segments are installed


under the protection of the shield, leaving a radial gap remains behind
the rear shield. Meanwhile, backlling with pea gravel followed by
grouting is performed at a certain distance behind the shield with a
consequence that an unsupported span exists. As shown in Fig. 23b,
the rock mass experiences three unloading processes during the excavation. Firstly, as the tunnel remains unsupported at 2 m behind the
face (Figure 17), the tunnel boundary experiences the rst unloading
process. Then, as soon as the gap is closed, loading of the shield
takes place. Secondly, the entire tunnel boundary is unloaded at the

rear shield due to the conicity of the machine. Until the pea gravel is
backlled at certain distance behind the shield, the last unloading
process occurs.
Consequently, as depicted in Fig. 24, a signicant stress concentration develops at the end of the front shield where contact with the
ground occurs, causing permanent damage to the TBM.
Moreover, the maximum stresses in tension in the lining can
also be checked in Fig. 25. This value (16 MPa) is in good agreement
with the monitoring data (15 MPa) of the extensometers installed
in the lining.

Fig. 21. Maximum principal stress contours around the excavation.

K. Zhao et al. / Engineering Geology 181 (2014) 93111

Fig. 22. Horizontal displacements on the left side of the tunnel.

Fig. 23. Results of the model: (a) Longitudinal displacement prole (LDP); (b) contact pressure developing on the shield and on the lining.

109

110

K. Zhao et al. / Engineering Geology 181 (2014) 93111

Fig. 24. Maximum stress in the shield and cutterhead.

5.3. Load transfer along the longitudinal direction


The key to understand this special failure mode is that it represents a
complex load transfer mechanism along the longitudinal direction,
resulting from the discontinuous behaviour of the rock mass (with a
fault zone) and the interaction between the ground, the TBM and the
support system. Indeed, the similar mechanism has been reported and
well explained through the evidence from geodetic monitoring at the
Messochora tunnel and the Tymfristos tunnel, Greece (Kontogianni
et al., 2004, 2008; Stiros and Kontogianni, 2009).
Combining the support of the core ahead of the face, the shield (after
closing the gap) and the backlled segmental lining, three arching
actions typically develop between each other, which determine the
load transfer in the longitudinal direction. However, the fault zone
destroys these arching effects due to its near-parallel orientation to
the tunnel axis. In that, the load transfer along the longitudinal direction

deteriorates, signicantly. High convergence, as a result, occurred at the


certain area associated with re-acceleration in the strain accumulation
in the neighbouring sections.
As the convergence closes the gap progressively, the shield starts to
support the ground and undertakes the role of the core after excavation
to stabilize the tunnel. The spatial stress redistribution due to the excavation front advance has minor effect on the previously excavated, deformed and stabilized tunnel segmental linings at large distances from
the front. Nevertheless, as the TBM advances leaving an unsupported
span behind the machine, a signicant secondary stress re-distribution
developed and transferred at bi-lateral directions. It would lead to a
heavy load concentration on the shield and also induce a further disturbance to trigger failure of the pillar rock between the wall and the fault.
As soon as the failure of the stiffer pillar rocks occurs, the stresses have
to be redistributed, causing a load increase in the weak material, and
also in previously less loaded stiff blocks. It transferred at distance of

Fig. 25. Maximum stress in the lining.

K. Zhao et al. / Engineering Geology 181 (2014) 93111

12 m (more than 2D) backwards along the tunnel, in areas remote from
the working face and presumably free from the face effect.
Through the results, it suggests that despite the extremely weak rock
conditions and rock yielding in the failure zone, ground improvement
(i.e., grouting and rock bolts) contributed to the development of the
physical bearing capacity of the rock around the excavation and also
to the improvement of the groundsupport interaction which sustained
stability around the opening.
Furthermore, the arching action between the core and segmental
lining leads to an additional ground deformations in this area and further increases the ground pressure acting on the shield. Therefore, the
stiffer the lining and the shorter its distance from the face, the more pronounced will be the arching effect and the less will be the ground pressure (Ramoni and Anagnostou, 2011). In this respect, backlling with
continuous grouting is more advantageous than pea gravel and mortar
with an unsupported span behind the shield (for the machine with a
given length). On the other hand, a stiff support with embedment
right behind the machine is favourable for the shield but, inevitably, attracts a higher ground load.
6. Conclusions
The brittle fault zone sub-parallel to the tunnel axis presented many
challenges to the engineers and contractors during the construction of
the Aicha exploratory tunnel (along the Brenner Base Tunnel). Serious
local instabilities occurred at the side wall (close to the fault zone)
during TBM drive. The segmental lining was collapsed at a distance of
more than 2D (D is tunnel diameter) behind the face, without any
evidence. The correct anticipation of the fault zones as the tunnel
advances and a timely increase both in reinforcement (grouted bolts)
and holding elements (strong anchored bolts and heavy mesh or
mesh over steel sets) are the key to success in this case.
A 3D numerical model is presented and applied in this paper, to
simulate this instability problem and in particular the specic failure
mode, i.e., the combination of the brittle failure of the pillar rock between the tunnel wall and the fault, and the shear yielding of the
gouge. It is seen that model is highly effective in reproducing the discontinuous behaviour of the rock mass (with a fault zone) and the interaction between the ground, the TBM and the support system.
With this further extension, the 3D numerical model developed
by the authors is capable to deal with the main mechanical adverse
conditions for TBM tunnelling in mountain terrain (as for example illustrated by Verman et al., 2012): brittle failure (spalling), squeezing and
fault zones. Nevertheless, the main problem which still remains is
the geological fault zones with the so-called swimming gouge (i.e.,
practically cohesionless material under high water pressure). It is
expected that a coupled hydraulic and mechanical analysis will be introduced in the near future.
Acknowledgment
The work of the present article was developed within the framework
of the Youth Fund of the Natural Science Foundation of Jiangsu Province
(no. BK 20130933) and the National Key Basic Research Program of

111

China (no. 2011CB013605). The authors would like to thank for the
precious nancial and moral support.
References
Aydin, A., Ozbek, A., Cobanoglu, I., 2004. Tunneling in difcult ground: a case study from
Dranaz tunnel, Sinop, Turkey. Eng. Geol. 74, 293301.
Barla, G., Ceriani, S., Fasanella, M., Lombardi, A., Malucelli, G., Martinotti, G., Oliva, F.,
Perello, P., Pizzarotti, E.M., Polazzo, F., Rabagliati, U., Skuk, S., Zurlo, R., 2010. Problemi
di stabilit al fronte durante lo scavo del cunicolo esplorativo AicaMules della Galleria di Base del Brennero. MIR 2010, Torino.
Cai, M., Kaiser, P.K., Tasaka, Y., Maejima, T., Morioka, H., Minami, M., 2004. Generalized
crack initiation and crack damage stress thresholds of brittle rock masses near underground excavations. Int. J. Rock Mech. Min. Sci. 41, 833847.
Dalgic, S., 2000. The inuence of weak rocks on excavation and support of the Beykoz
Tunnel, Turkey. Eng. Geol. 58, 137148.
Dalgic, S., 2003. Tunneling in squeezing rock, the Bolu tunnel, Anatolian Motorway,
Turkey. Eng. Geol. 67, 7396.
Diederichs, M.S., 2003. Manuel Rocha Medal Recipient, rock fracture and collapse under
low connement condition. Rock Mech. Rock. Eng. 36 (5), 339381.
Diederichs, M.S., 2007. The 2003 Canadian Geotechnical Colloquium: mechanistic interpretation and practical application of damage and spalling prediction criteria for
deep tunneling. Can. Geotech. J. 44, 10821116.
European Commission, Mobility & Transport, TEN-T/Transport infrastructures, 2011. List
of the pre-identied projects on the core network in the eld of transport. http://
ec.europa.eu/transport/infrastructure/connecting/doc/revision/list-of-projects-cef.
pdf.
Grandori, R., Beniawski, Z.T., Vizzino, D., Lizzadro, L., Romualdi, P., Busillo, A., 2011. Hard
rock extreme conditions in the rst 10 km of TBM driven Brenner Exploratory
Tunnel. Proc. of Rapid excavation & Tunnelling Conference 2011, San Francisco, USA.
Hoek, E., Guevara, R., 2009. Overcoming squeezing in the YacambQuibor Tunnel,
Venezuela. Rock Mech. Rock. Eng. 42 (2), 389418.
Hoek, E., Marinos, P., 2010. Tunnelling in overstressed rocks. Rock engineering in difcult
ground conditions soft rocks and karst. Taylor & Francis Group, London.
Hoek, E., Marinos, P., Benissi, M., 1998. Applicability of the Geological Strength Index (GSI)
classication for very weak and sheared rock masses. The case of the Athens Schist
Formation. Bull. Eng. Geol. Environ. 57, 151160.
Kaiser, P.K., Diederichs, M.S., Martin, C.D., Sharp, J., Steiner, W., 2000. Invited keynote:
underground works in hard rock tunneling and mining. Geotech. Eng.Technomic
Publishing, Melbourne, Pennsylvania (USA), pp. 841937.
Kontogianni, V., Stiros, S., 2005. Induced deformation during tunnel monitoring: evidence
from geodetic monitoring. Eng. Geol. 79, 115126.
Kontogianni, V., Stiros, S., 2006. Convergence in Weak Rock Tunnels. 2. Tunnels and
Tunneling International, pp. 3639.
Kontogianni, V., Tzortzis, A., Stiros, S., 2004. Deformation and failure of the Tymfristos
tunnel, Greece. J. Geotech. Geoenviron. Eng. ASCE 130, 10041013.
Kontogianni, V., Papantonopoulos, Stiros, S., 2008. Delayed failure at the Messochora
tunnel, Greece. Tunn. Undergr. Space Technol. 23, 232240.
Kovari, K., Fechtig, R., 2000. Historical Tunnels in the Swiss Alps. Staubli AG, Zrich.
Loew, S., Barla, G., Diederichs, M., 2010. Engineering Geology of Alpine tunnels: past,
present and future. 11th IAEG Congress, Auckland, New Zealand.
Marinos, V., Fortsakis, P., Prountzopoulos, G., 2006. Estimation of rock mass properties of
heavily sheared ysch using data from tunneling construction. IAEG2006, Nottingham, United Kingdom.
Midas, 2010. Geotechnical and Tunnel Analysis System, User's Guide.
Ramoni, M., Anagnostou, G., 2011. The interaction between shield, ground and tunnel
support in TBM tunnelling through squeezing conditions. Rock Mech. Rock. Eng. 44,
3761.
Stiros, S.C., Kontogianni, V.A., 2009. Coulomb stress changes: from earthquakes to
underground excavation failures. Int. J. Rock Mech. Min. Sci. 46, 182187.
Verman, M., Carter, T.G., Babenderde, L., 2012. TBM vs D&B a difcult choice in mountain terrain some geotechnical guidelines. In: Harmonising Rock Engineering and
the Environment Qian & Zhou (Ed.), Taylor & Francis Group, London.
Vermeer, P.A., De Borst, R., 1984. Non-associated plasticity for soils, concrete and rock.
Heron 29, 164.
Vlachopoulos, N., Diederichs, M.S., 2009. Improved longitudinal displacements proles for
convergence connement analysis of deep tunnel. Rock Mech. Rock. Eng. 42,
131146.
Zhao, K., Janutolo, M., Barla, G., 2012. A completely 3D model for the simulation of
mechanized tunnel excavation. Rock Mech. Rock. Eng. 45 (4), 475497.

Das könnte Ihnen auch gefallen