Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Engineering Geology 265 (2020) 105405

Contents lists available at ScienceDirect

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

Assessment of non-linear rock strength parameters for the estimation of T


pipe-jacking forces. Part 2. Numerical modeling

C.S. Chooa, , D.E.L. Ongb
a
Research Centre for Sustainable Technologies, Faculty of Engineering, Computing & Science, Swinburne University of Technology Sarawak Campus, 93350 Kuching,
Sarawak, Malaysia
b
School of Engineering and Built Environment, Griffith University, 170 Kessels Rd., Nathan, Queensland 4111, Australia

A R T I C LE I N FO A B S T R A C T

Keywords: In the earlier companion paper (Ong and Choo, 2018), direct shear tests were performed on reconstituted
Pipe-jacking tunneling rock spoils for the development of equivalent tangential cohesion, c't,p and friction angle, ϕ't,p. These
Friction equivalent rock strength parameters were used in assessing jacking forces, which were measured from three
Finite element pipe-jacking drives negotiating the highly weathered lithologies from the Tuang Formation, Malaysia. The
Arching ratio
measured jacking forces were back-analyzed, leading to the development of frictional coefficient, μavg and the
vertical stress at the tunnel crown, σEV. In this paper, the reliability of using the parameters, c't,p, ϕ't,p, μavg and σEV
was assessed through the use of three-dimensional finite element modeling of the studied pipe-jacking drives.
The peak tangential strength parameters, c't,p and ϕ't,p were applied to the modeled rock mass and the strength
parameters of the pipe-rock interface were developed from μavg. Arching ratios were developed from σEV, which
were subsequently used for post-analysis assessment of the results from numerical modeling. There was sound
agreement between the resulting pipe-rock interface shear stresses, τ1 and reaction loads, ΣFy from the numerical
analysis against the field measured jacking forces, thus demonstrating the reliability of the use of developed
equivalent rock strength parameters, c't,p, ϕ't,p, μavg and σEV in the assessment of pipe-jacking forces through
highly weathered lithologies.

1. Introduction From the backanalysis of the jacking forces, the developed average
frictional coefficient, μavg were used to indicate the nature of pipe-rock
In the earlier companion paper (Ong and Choo, 2018), tunneling friction, i.e. lubricated or unlubricated. The stresses acting normal to
rock spoils were used for the backanalysis of jacking forces. This the outer periphery of the pipe crown, σEV was also developed from the
method has been used successfully for pipe-jacking drives traversing tangential strength parameters. Values of σEV were used to indicate the
lithological units of sandstone, shale and phyllite origins (Choo and degree of arching, as arching effect was found to be rather significant in
Ong, 2015). From the earlier companion paper, direct shear tests were the assessment of pipe-jacking forces.
performed on reconstituted tunneling rock spoils of shale and inter- The work described in the current paper was motivated by the need
bedded lithological units of metagraywacke – siltstone and greywacke – to evaluate the reliability of the peak tangential strength parameters,
phyllite origins. It was essential for the direct shear test results to be c't,p and ϕ't,p (developed from direct shear testing) and μavg and σEV
relevant in the understanding of jacking forces for pipe-jacking drives (developed from backanalysis of field measured jacking forces), parti-
through the highly weathered and fractured Tuang Formation in cularly for the assessment of pipe-jacking forces in highly weathered
Kuching City, Malaysia. This necessitated the development of tangential and fractured lithologies. Hence, the reliability of using these para-
strength parameters, c't,p and ϕ't,p based on arching theory. Typically, meters was assessed through three-dimensional (3D) finite element
arching theory is more applicable to soil. However, in highly fractured modeling of the pipe-jacking drives studied in this paper. The peak
litohologies, it is possible for the highly weathered rock to demonstrate tangential strength properties, c't,p and ϕ't,p were applied to the modeled
soil-like behavior, akin to a highly weathered ‘soft rock’ (Choo and Ong, rock mass, while the values of backanalyzed μavg were applied to the
2015). In the current study, the consideration of the traversed rock strength models of the pipe-rock interface elements. Prescribed dis-
masses as ‘soft rock’ will be explained, hereinafter. placements were used to simulate the forward advancement of the rigid


Corresponding author.
E-mail addresses: cschoo@swinburne.edu.my (C.S. Choo), d.ong@griffith.edu.au (D.E.L. Ong).

https://doi.org/10.1016/j.enggeo.2019.105405
Received 31 July 2018; Received in revised form 5 November 2019; Accepted 7 November 2019
Available online 15 November 2019
0013-7952/ © 2019 Elsevier B.V. All rights reserved.
C.S. Choo and D.E.L. Ong Engineering Geology 265 (2020) 105405

‘wish-in-place’ concrete pipeline. This modeling technique was vali- Table 1


dated through the use of the lower bound theorem for assessing the Parameters used in pipe-jacking force model for backanalyses of μavg (Ong and
incurred pipe-rock interface shear stresses, τ1 and the reaction loads, Choo, 2018).
ΣFy. For the consideration of arching in the results from the numerical Drive Drive A Drive B Drive C
analysis of the pipe-jacking drives, arching ratios were developed from
the backanalyzed σEV, and subsequently applied to jacking forces Geology Shale Metagraywacke – Graywacke –
siltstone phyllite
computed from τ1 and ΣFy as post numerical analysis.
Length of rock cores extracted from 6.0 12.0 13.0
shaft locations (m)
2. Numerical modeling of tunnels RQD (%) Maximum 23 20 20
Minimum 0 0 0
Average 14.0 5.8 2.0
Several studies have used continuum methods in the modeling of
Length of rock cores with RQD = 0 0.7 8.9 11.0
tunnels negotiating rock masses (Basarir et al., 2005; Ding and Li, 2011; (m)
Li et al., 2018, 2019; Pusch and Börgesson, 1998; Valliappan and Ang, External pipe diameter, De (m) 1.78 1.43 1.78
1988; Zhang et al., 2006; Zhu et al., 2003). Basarir et al. (2005)) and Soil unit weight, γsoil (kN/m3) 18 18 18
Genis et al. (2007)) utilized continuum-based finite element analysis for Depth to rock level, hsoil (m) 20 9 17
Rock unit weight, γrock (kN/m3) 22 22 22
the simulation of tunnels traversing fractured rock masses from Turkey
Height of rock cover, hrock (m) 1 11 10
for the purpose of designing tunnel support systems. These rock masses Tangential MC c't,p (kN/m2) 50.8 57.8 29.0
included moderately weathered limestone, highly weathered diabase parameters ϕ't,p (°) 47.8 44.3 38.7
(with average RQD of 28%), and moderately weathered sandstone obtained from
direct shear
(with average RQD of 34%) (Basarir et al., 2005), as well as heavily
testing
broken phyllite and moderately weathered granodiorite (Genis et al., Average measured jacking forces, 37.2 20.7 23.6
2007). Pusch and Börgesson (1998) proposed functions for the com- JFmeas (kN/m)
putation of rock mass moduli and stiffnesses of discontinuities through Vertical stress acting on pipe crown, 12.5 −19.9 −38.5
the use of finite element methods. Zhu et al. (2003) simulated tunnel σEV (kN/m2) (see Eq. (3))
Back-analyzed μavg, using Eq. (1) 0.356 0.491 0.581
excavation schemes (single- and multi-benching) for various rock con-
R2 0.86 0.70 0.77
ditions (including soft to hard rocks as well as fault materials) in
Taiwan in order to study tunnel stability through the use finite element
method. The continuum-based models were able to simulate the de- significant factor in the understanding of jacking forces. In a subsequent
velopment of plastic zones in the rock masses around the tunnel, thus study, Yen and Shou (2015) utilized the displacement control approach
allowing the study of the distribution of rock movements around the to simulate the jacking of pipes. Other studies have also showed that the
tunnel. These past studies show that continuum-based finite element forward advancement of the pipeline could be simulated through the
methods can be used for the modeling of tunnels traversing rock masses use of prescribed displacements on the pipeline elements (Dounias and
of assorted lithologies with varying weathering conditions. Potts, 1993; Hu and Pu, 2004; Merifield and Sloan, 2006).
However, the numerical simulation of pipe-jacking drives is less The aforementioned studies on numerical modeling emphasize the
common, particularly when modeling the advance of the pipeline use of continuum-based finite element analysis in the simulation of
through a rock mass. Li et al. (2019) used continuum-based finite ele- pipe-jacking drives through soils and fractured rocks. The use of pre-
ment analysis software to study the interfacial contact area between scribed displacements is also necessary in the modeling of a rigid
rocks (with behavior simulated using the Mohr-Coulomb failure cri- concrete pipeline, traversing a model rock mass. The findings from the
terion) and pipes during pipe-jacking. The continuum models com- earlier studies will form the basis for the numerical analysis of pipe-
prised of contact conditions along the modeled pipeline. This model jacking forces conducted, hereinafter.
was used to simulate rock-pipe interfacial stresses from a case study of a
pipe-jacking drive traversing lithological units of sandstone. Barla et al.
(2006) performed 2D analyses of pipe-jacking drives through fractured 3. Numerical analysis of pipe-jacking forces
limestone, using continuum and discontinuum modeling techniques.
The initial continuum modeling method was performed using FLAC 3D finite element analysis was performed on the pipe-jacking drives
(Itasca Consulting Group Inc, 2001). Through the reduction of tensile in this paper through the use of PLAXIS 3D, a commercially available
strength of the rock mass, the modeled rock mass failed in a localized finite element software package for geotechnical engineering purposes.
zone at the tunnel crown, with lower tensile strength resulting in larger The results from the direct shear testing of reconstituted tunneling rock
failure zones. In order to study the influence of discontinuities in the spoils and the subsequent backanalysis of jacking forces are shown in
rock mass on pipe-jacking forces, a discontinuum analysis was per- Table 1, along with properties related to the traversed geology.
formed using UDEC (Itasca Consulting Group Inc, 1999). The Geolo-
gical Strength Index (GSI) of the modeled rock was varied. The nu-
merical simulation found that the discontinuities in the rock mass 3.1. Finite element mesh
resulted in high jacking forces, which led to a stoppage in pipe-jacking
works. The work by Li et al. (2019) and Barla et al. (2006) demon- The finite element models used in this study measured 30 m wide in
strated the possibility of modeling pipe-jacking drives in fractured the x-direction (transverse to pipeline), 20 m long in the y-direction
lithologies. (longitudinal to pipeline), and 30 m deep in the z-direction. Fig. 1
Shou et al. (2010) modeled linear and curved pipe-jacking align- shows the typical finite element mesh used for the numerical analysis
ments in 3D space to assess the effect of lubrication in the overcut hereinafter. 10-node tetrahedral elements were used to simulate the
annulus. The simulated pipe-jacking drives were jacked through the rock and soil masses, while 6-node plate elements were used to model
model soil based on repeated sequence of deactivating soil elements at the concrete pipelines. For the modeling of interactions between the
the tunnel face, followed by a forward displacement of the modeled driven pipelines and the surrounding rock mass, 12-node interface
pipeline. The effect of lubrication was simulated by varying the pipe- elements were placed along the outer periphery of the driven pipelines.
soil frictional coefficient, f. It was found that jacking forces were not A set of general fixities were automatically generated and imposed
only a consequence of lubrication, but that the contact area between the along the boundaries of the model. The boundary conditions are as
outer surface of the pipeline and the surrounding soil mass was a follows:

2
C.S. Choo and D.E.L. Ong Engineering Geology 265 (2020) 105405

assumed for the overburden soil and rock mass surrounding the tunnel
(Choo and Ong, 2015).
The stiffnesses of the modeled rock masses were developed from the
observations of RQD values being predominantly zero for typically
highly weathered and fractured rock masses, such as the Tuang
Formation. This allowed for the estimation of rock mass stiffness using
the following equations by Bieniawski (1989).
EM = 10(RMR − 10)/40 (3)

RMR = 9 ln Q + 44 (4)

RQD J J
Q= × r × w
Jn Ja SRF (5)
where EM is the stiffness of the rock mass; RMR is the rock mass rating
(Bieniawski, 1979); and Q is the rock mass quality. Q is a function of
various rock joint properties including RQD (Barton et al., 1974), Jn
which is the joint set number; Jr which is the joint roughness number; Ja
which is the joint alteration number; Jw which is the joint water re-
Fig. 1. Typical model used for numerical analysis of interface shear stresses. duction number; and SRF which is the stress reduction factor (Barton
et al., 1974).

• Rollers were applied to vertical boundary faces normal to the x- and


For estimating Q, the values for relevant parameters were difficult
to determine due to the absence of any exposed rock outcrops, or rock
y-directions, restricting horizontal movements in the x- and y-di-
faces during shaft construction. These exposed rock faces were quickly
rections respectively, while allowing for vertical movements;

shotcreted in order to mitigate water ingress, which would have
Full fixities were applied to the bottom boundary, restricting
otherwise caused ground surface deformation (Choo and Ong, 2015).
movements in all directions;

This prevented joint directions from being thoroughly studied. With
The ground surface (top boundary) was free to move in all direc-
this limited information, assumptions had to be made in estimating the
tions.
Q rating, and subsequently the stiffness of the rock mass, EM. These
assumptions are shown in Table 2. The resulting rock modulus value of
3.2. Rock and soil mass properties 7.1 GPa was applied to the numerical modeling of the highly fractured
lithologies with homogenously and consistently very poor quality as
Rock masses for the studied drives were modeled using the elastic- described in the Q system (Bieniawski, 1989). The overburden rock
perfectly plastic Mohr-Coulomb (MC) failure criterion, together with mass and soil mass properties are shown in Tables 3, 4 and 5 for Drives
the peak tangential strength parameters, c't,p and ϕ't,p developed in the A, B and C, respectively.
earlier companion paper (Ong and Choo, 2018). The MC failure cri-
terion was used to simulate the constitutive behavior of the rock mass. 3.3. Pipeline and rock-pipe interface properties
These MC strength parameters are used in the Pellet-Beaucour and
Kastner (2002) jacking force model, which is given as The concrete pipelines were modeled as 20 m long hollow cylinders,
π γDe γDe ⎤ longitudinal in the y-direction. The geometry of the modeled pipeline
F = μLDe ⎡ ⎛σEV + ⎞ + K2 ⎛σEV + ⎞ was created by initially generating a circular polycurve of a diameter
2⎣⎢⎝ 2 ⎠ ⎝ 2 ⎠⎦⎥ (1)
appropriate to the respective pipe-jacking drives. The diameters of the
modeled pipelines are shown in Tables 3, 4 and 5 for Drives A, B and C,
(
b γ−
2C
)
σEV =
2K tan ϕ
b
(1 − e −2K h tan ϕ
b
) (2)
respectively. The polycurve was positioned on a vertical boundary. The
polycurve was subsequently extruded 20 m in the y-direction, creating a
hollow cylinder which had both ends flushed with vertical model
The Pellet-Beaucour and Kastner jacking force model was developed
boundaries. No tunnel face was considered in this simulation since the
for drives traversing sands and clays (Pellet-Beaucour and Kastner,
earlier study (Ong and Choo, 2018) was focused on studying frictional
2002). It considers soil arching phenomenon; therefore, it is dependent
jacking forces. This was reasonable since frictional jacking forces in-
on MC soil strength parameters. This would be applicable to the ‘soft
crease with drive span, and are usually the major component in jacking
rock’ encountered in the current study, justifying the need to develop
forces.
strength parameters c't,p and ϕ't,p from the direct shear tests performed
The rock-pipe interface was modeled using 12-node interface ele-
earlier (Ong and Choo, 2018). This would allow for an assessment of the
ments. The interface elements are parings of nodes, thus ensuring
use of the tangential strength parameters in numerical modeling of the
compatibility with the 6-noded triangular faces of both soil and plate
surrounding rock mass, as well as in the backanalysis of jacking forces,
elements. The use of finite elements for simulating the behavior of
which was performed earlier in Ong and Choo (2018). Peak strength
properties were used for both backanalysis and numerical modeling due
Table 2
to the consideration of arching in the Pellet-Beaucour and Kastner Parameters used for estimating Q rating (Bieniawski, 1989).
(2002) jacking force model. During the reduction in soil stresses due to
tunneling, arching is initiated by a vertical slippage of the overburden Parameter Description Value

rock or soil mass. The initial vertical slip of rock or soil into the ex- Rock quality designation, RQD Very poor 0–25
cavated tunnel occurs at low displacements. Hence, the use of peak Joint set number, Jn Random, heavily jointed 15
strength parameters, c't,p and ϕ't,p for the rock mass was justified. The Joint roughness number, Jr Rough 1.5
Pellet-Beaucour and Kastner (2002) model assumed uniform isotropic Joint alteration number, Ja Unaltered joint walls 1.0
Joint water reduction factor, Medium inflow or pressure occasional 1.0
conditions for the backanalysis of jacking forces, i.e. K0 = 1. Therefore,
Jw outwash of joint fillings
in order to further facilitate parity between the backanalysis of jacking Stress reduction factor, SRF Medium stress 1.0
forces and the numerical analysis hereinafter, isotropic conditions were

3
C.S. Choo and D.E.L. Ong Engineering Geology 265 (2020) 105405

Table 3
Material properties for finite element modeling of Drive A.
Material Loose sand Very soft Loose sand Dense sand Medium dense Very soft Very hard Shale Rock-pipe Pipe
clay sand clay silt interface

Elements used Soil Soil Soil Soil Soil Soil Soil Soil Interface Plate
Material model MC MC MC MC MC MC MC MC MC Linear elastic
Depth (m) 0 to 6 6 to 10.5 10.5 to 11.5 to 13.5 to 14.9 14.9 to 19.3 to 20 20 to 30 – –
11.5 13.5 19.3
Unit weight, γ (kN/m3) 18 18 18 18 18 18 18 22 – 24
SPT ‘N’ 6 2 6 34 19 2 107 – – –
Cohesion, c' (kPa) 5 5 5 15 10 5 15 21.7(1) 0 –
Friction angle, ϕ' (°) 28 28 28 30 28 28 33 38.7(1) 19.60(2) –
Elastic modulus, E' (kN/ 10,430 2200 10,430 59,130 33,040 2200 117,700 7.1 × 106 1000 27 × 109
m2)
Poisson's ratio, ν 0.35 0.35 0.35 0.35 0.35 0.35 0.35 0.35 0.35 0.15
External pipe diameter, De – – – – – – – – – 1.78
(m)

Note: 1. From generalized tangential interpretation of earlier direct shear tests.


2. Interface friction angle, ϕi = tan−1(μavg).

highly weathered lithologies has been mentioned previously, and has friction angle; and μavg is the back-analyzed average frictional coeffi-
been used successfully when applied to tunnels in highly weathered cient, obtained from earlier back-analyses of jacking forces (Ong and
rocks (Basarir et al., 2005; Genis et al., 2007; Li et al., 2019; Ong and Choo, 2018).
Choo, 2016; Pusch and Börgesson, 1998; Zhu et al., 2003). To simulate In addition to the treatment of concrete pipelines as rigid bodies and
the longitudinal displacement of the concrete pipeline in the bored the use of user-defined prescribed displacements, it was necessary to
tunnel, it was necessary to nullify the effects of the boundary fixities at assign suitable values to the stiffness of the pipe-rock interface ele-
the pipe ends. This was achieved through the use of user-defined pre- ments, to facilitate the segregation of the pipe and rock elements, thus
scribed displacements at the ends of the pipelines. The application of allowing for longitudinal displacement of the concrete pipeline. This
user-defined prescribed displacements at any point along the bound- was achieved by prescribing a low value to interface modulus, Einter.
aries prevail the effects of any automatically generated boundary fix- The use of low interface moduli has been applied to simulations of
ities (Plaxis bv, 2013). Prescribed displacements allowed for the mea- slippage between two materials. Yin et al. (1995) and Wang and Wang,
surement of reaction forces, ΣFy, the use of which will be explained, 2006 adopted low tangential interface modulus values to enable large
hereinafter. interface deformations for studies on pull-out tests. For the current
The modeled concrete pipelines were regarded as rigid bodies to study on modeling the large deformation of pipe-rock interfaces, an
allow for the segregation of compressive axial forces in the modeled interface modulus of Einter of 1000 kPa was used as it was still sig-
pipelines from the pipe-rock shear stresses generated at the rock-pipe nificantly small relative to the moduli of the concrete pipeline and
interfaces. The treatment of the modeled pipelines as rigid bodies was surrounding rock mass. This method of modeling the rock-pipe inter-
justifiable since the concrete pipelines were of a higher stiffness in re- face was also used successfully by Ong and Choo (2016).
lation to the less rigid pipe-rock interface surrounding the modeled
pipelines (Tian et al., 2014; Wang et al., 2009; Zhang et al., 2013). 3.4. Calculation phases in modeling of pipe-jacking process
The pipe-rock interface was simulated using the Mohr-Coulomb
model, which allowed for direct prescription of rock-pipe interface For simulating the propulsion of the concrete pipeline through the
strength properties from the earlier back-analyses of frictional jacking rock mass, modeling of the pipe-jacking process was performed in three
forces (Ong and Choo, 2018). The interfacial strength properties are consecutive phases.
given as
τi = σ ′n tan ϕi • Stage 1 – Application of initial stress conditions
• Stage 2 – Wish-in-place pipeline and activation of rock-pipe inter-
(6)

ϕi = tan−1 (μavg ) face elements


• Stage 3 – Application of displacements at tail-end of pipeline
(7)

where, τi is the rock-pipe interface shear strength; σ'n is the effective


normal stress acting on the rock-pipe interface; ϕi is the interface In Stage 1, material properties for the soil and rock mases,

Table 4
Material properties for finite element modeling of Drive B.
Material Soft clay Soft silt Very stiff silt Hard silt Meta-graywacke – siltstone Rock-pipe interface Pipe

Elements used Soil Soil Soil Soil Soil Interface Plate


Material model MC MC MC MC MC MC Linear elastic
Depth (m) 0 to 2.7 2.7 to 5.5 5.5 to 7 7 to 9 9 to 25 – –
Unit weight, γ (kN/m3) 18 18 18 18 22 – 24
SPT ‘N’ 3 3 29 54 – – –
Cohesion, c' (kPa) 5 5 10 15 44.6a 0 –
Friction angle, ϕ' (°) 28 28 28 33 39.9a 26.15b –
Elastic modulus, E' (kN/m2) 3300 3300 31,900 58,960 7.1 × 106 1000 27 × 109
Poisson's ratio, ν 0.35 0.35 0.35 0.35 0.35 0.35 0.15
External pipe diameter, De (m) – – – – – – 1.43

Note: a. From generalized tangential interpretation of earlier direct shear tests.


b. Interface friction angle, ϕi = tan−1(μavg).

4
C.S. Choo and D.E.L. Ong Engineering Geology 265 (2020) 105405

Table 5
Material properties for finite element modeling of Drive C.
Material Loose sand Stiff silt Hard silt Hard silt Graywacke – phyllite Rock-pipe interface Pipe

Elements used Soil Soil Soil Soil Soil Interface Plate


Material model MC MC MC MC MC MC Linear elastic
Depth (m) 0 to 5 5 to 8 8 to 12.5 12.5 to 17 17 to 35 – –
Unit weight, γ (kN/m3) 18 18 18 18 22 – 24
SPT ‘N’ 3 9 35 66 – – –
Cohesion, c’ (kPa) 5 5 15 15 60.8a 0 –
Friction angle, ϕ’ (°) 28 28 30 33 41.8a 30.16b –
Elastic modulus, E' (kN/m2) 5217 9900 38,500 72,600 7.1 × 106 1000 27 × 109
Poisson's ratio, ν 0.35 0.35 0.35 0.35 0.35 0.35 0.15
External pipe diameter, De (m) – – – – – 1.43

Note: a. From generalized tangential interpretation of earlier direct shear tests.


b. Interface friction angle, ϕi = tan−1(μavg).

groundwater levels and K0 condition were assigned to the model for the resulting pipe-rock interface shear stresses. This equilibrium needs to be
generation of initial stresses under greenfield conditions. Model fulfilled at all calculation stages.
boundary conditions were also activated. Plate elements, interface For this validation, the lower bound theorem was applied to Drive
elements, prescribed displacements and prescribed loads were deacti- A, under which three cases were assessed:
vated in Stage 1. These elements and loads would only be activated in
subsequent stages. • Case 1 (parent): No displacement or loading was applied to the
Following the generation of initial stresses in Stage 1, the ‘wish-in- modeled pipe.
place' 20 m long concrete pipeline was simulated in Stage 2 through the • Case 2 (child): Displacement was prescribed to the modeled pipe.
activation of plate elements. This was accompanied with the deacti- • Case 3 (child): Loading was applied to the modeled pipe.
vation of soil elements and water conditions within the pipeline an-
nulus. Rock-pipe interface elements were also activated concurrently Case 1 was performed as a baseline for when displacements (Case 2)
with the plate elements. or loads (Case 3) were applied to the pipe. No interface shear stresses
In Stage 3, the user-defined displacements were prescribed hor- were expected to develop in Case 1. Case 2 was performed to verify that
izontally in the y-direction and applied along the periphery of the tail the modeling technique (including the earlier described elements,
end of the modeled pipeline. This was coupled with the application of parameters and values, and the use of prescribed displacements to
zero-valued prescribed displacements at the front end of the pipe to displace the modeled pipeline) was suitable for mobilizing interfacial
mitigate the effect of boundary fixities. These prescribed displacements shear stresses, as described in the aforementioned literature (Dounias
were free in the y-direction (longitudinal to the modeled pipeline), thus and Potts, 1993; Hu and Pu, 2004; Merifield and Sloan, 2006). Case 3
allowing for the movement of the modeled pipeline beyond the model was performed as a compliance check of Case 2, to verify that the nu-
boundaries (Plaxis bv, 2013). merical model was valid for the application of applied loads to displace
the modeled pipeline. The successful modeling of Cases 2 and 3 would
4. Validation of numerical analysis technique show that both prescribed displacements and applied loads were sui-
table in simulating the driving of the modeled pipelines. Cases 2 and 3
In order to assess the applicability of direct shear test results to the also served to demonstrate that the model was a lower bound solution
assessment of pipe-jacking forces, it was necessary to validate the ear- (Randolph and Houlsby, 1984).
lier described numerical analysis technique. The validation was also
necessary due to the limited use of prescribed displacements in mod- 4.1. Case 1 – No displacements or loading applied to modeled pipe
eling the advance of a pipeline during pipe-jacking (Barla et al., 2006;
Shou et al., 2010; Yen and Shou, 2015). This validation was performed In Case 1, displacements and loads were not applied to the modeled
by applying the lower bound theorem to three baseline cases. pipeline. This case was performed in order to establish a reference point
The lower bound theorem is widely used to verify that static equi- for Cases 2 and 3, where displacements and loads would be applied.
librium has been achieved in a system. This leads to a safe solution for Inspection of the pipe-rock interface elements showed that the elements
the simulated geotechnical problem. Some geotechnical problems were all in an elastic state (see Fig. 2). This was verified by the dis-
which have used the lower bound theorem include ground anchors tribution of constant interface shear stress, τ1 longitudinally along (see
(Merifield and Sloan, 2006), underground excavations (Augarde et al., Fig. 3) and radially around the pipeline (see Fig. 4). The pipe-rock in-
2003; Davis et al., 1980; Smith, 1998, 2012), and pile performance terface shear stresses were uniformly distributed at τ1 = 0. Minimal
(Finno et al., 1991; Georgiadis et al., 2013; Randolph and Houlsby, shear stresses were observed at the ends of the modeled pipeline. This
1984). According to Randolph and Houlsby (1984), a lower bound was due to end effects at the model boundaries.
solution is achieved when the following criteria have been met: With no displacements or loading applied, the pipe-rock interface
elements exhibited the anticipated behavior. The results from Case 1
• The lower bound on the applied loads will cause collapse of per- would form the baseline for subsequent validation of the modeling
fectly plastic material with an associated flow rule; technique under prescribed displacements (Case 2) and applied loads
• The stress field is in equilibrium with the applied loads and (Case 3).
• The stress field does not violate the yield criterion for the material.
4.2. Case 2 – Displacements prescribed to modeled pipe
In addition to these criteria, Potts and Zdravković (1999) stated that
such a lower bound solution should be found in a stress field that is Case 2 follows from the completion of Case 1. A uniform prescribed
statically admissible, i.e. a system in equilibrium. Applying these cri- displacement of 100 mm was applied along the pipe circumference at
teria to the earlier described numerical analysis technique would ne- the tail end of the modeled pipeline (see Fig. 5). The prescribed dis-
cessitate that the applied jacking loads were in equilibrium with the placement was applied in the horizontal y-direction to allow for the

5
C.S. Choo and D.E.L. Ong Engineering Geology 265 (2020) 105405

Fig. 5. Validation of numerical analysis: Case 2 – Displacements (indicated by


arrows) prescribed to tail end of modeled pipeline.

Fig. 2. Validation of numerical analysis: Case 1 – All rock-pipe interface ele-


ments exhibiting elasticity (indicated by octahedron symbols).

Fig. 3. Validation of numerical analysis: Case 1 – Distribution of rock-pipe in-


terface shear stresses, τ1 along longitudinal cross-section of modeled pipeline.

Fig. 6. Validation of numerical analysis: Case 2 – All rock-pipe interface ele-


ments exhibiting plasticity (indicated by cube symbols).

modeled pipeline of a uniform diameter would exhibit uniform dis-


tributions of longitudinal and radial pipe-rock interface shear stresses,
τ1.
After the application of prescribed displacements, the rock-pipe
Fig. 4. Validation of numerical analysis: Case 1 – Radial distribution of rock- interface shear stresses, τ1 were uniformly distributed radially (see
pipe interface shear stresses, τ1 at mid-span of modeled pipeline. Fig. 7) and longitudinally (see Fig. 8) at τ1 = 65 kN/m2. From Eqs. (6)
and (7), an assessment of τ1 was performed. Based on σ'n = 172 kN/m2
and μavg = 0.356, the resulting interface shear strength, τi was 61 kN/
modeled pipeline to displace as a rigid body longitudinally. The pre-
m2, which was approximately equal to the resulting interface shear
scribed displacement of 100 mm was sufficient for the rock-pipe in-
stresses, τ1 = 65 kN/m2 from numerical analysis. This reconfirmed the
terface elements to reach a state of elastic yield, as shown by the pipe-
earlier observations from Fig. 6 that the rock-pipe interface elements
rock interface elements achieving plasticity (see Fig. 6). During the
had yielded.
backanalysis of jacking forces using the Pellet-Beaucour and Kastner
The reaction force at yield in the y-direction, ΣFy,yield was 7233 kN.
(2002) jacking force model, the rate of increase in jacking forces re-
By integrating τ1 along the outer surface area of the pipe, the total
mained uniform under homogenous geological conditions. Hence, a

6
C.S. Choo and D.E.L. Ong Engineering Geology 265 (2020) 105405

Fig. 9. Validation of numerical analysis: Case 3 – Loads applied to tail end of


modeled pipeline.

4.3. Case 3 – Loading applied to modeled pipe

Case 3 followed from the completion of Case 1. In contrast with Case


2 where prescribed displacements were used, Case 3 involved the use of
Fig. 7. Validation of numerical analysis: Case 2 – Uniform radial distribution of
applied loads to propel the modeled pipeline forward. Case 3 was
rock-pipe interface shear stresses, τ1 (65 kN/m2) at mid-span of modeled pi-
performed as a compliance model to show that static equilibrium of the
peline.
system could also be achieved through the use of applied loads.
Successful modeling of Case 3 would also justify the use of reaction
force, ΣFy for the assessment of jacking forces. From the results in Case
2, the reaction force at yield was ΣFy,yield = 7233 kN. Hence, in order to
ensure model convergence, the applied load, Fy,app in Case 3 could not
exceed 7233 kN, i.e. Fy,app < ΣFy,yield. A uniform line load of 100 kN/m
Fig. 8. Validation of numerical analysis: Case 2 – Uniform distribution of rock- was applied along the circumference at the tail end of the pipe, corre-
pipe interface shear stresses, τ1 (65 kN/m2) along longitudinal cross-section of sponding with a total applied load, Fy,app of 559 kN (see Fig. 9). As the
modeled pipeline. Pellet-Beaucour jacking force model is based on the strength of the
pipe-rock interface, sufficient displacement was determined by the at-
jacking force due to pipe-rock interface shear stresses, Fτ,cal can be given tainment of elastic yield. Therefore, it was necessary to show that under
as the prescribed displacement, a uniform distribution of interface shear
stresses was attained. The distribution of interface shear stresses long-
Fτ , cal = τ1 πDe L (8) itudinally (see Fig. 10) and circumferentially (see Fig. 11) were uniform
at τ1 = 5 kN/m2. Using Eqs. (6) and (7), integrating the pipe-rock in-
where, L is the length of the modeled pipeline; and De is the external terface shear stresses along the modeled pipeline resulted in a total
diameter of the pipe, which was 1.78 m for Drive A. This resulted in jacking force, Fτ,cal of 559 kN, which was in equilibrium with Fy,app. The
Fτ,cal = 7270 kN, which was comparable to the yield reaction force, achievement of static equilibrium in Case 3. between the applied loads,
ΣFy,yield of 7233 kN. The slight discrepancy was attributed to the end the reaction loads and the pipe-rock interface shear stresses justified the
effects (see Fig. 8), since τ1 was measured at the mid-span of the use of ΣFy in assessing pipe-jacking forces.
modeled pipeline (see Fig. 7). Despite the minor discrepancy, it can be
seen that the reaction forces due to the prescribed displacements were
equivalent to the total jacking forces due to pipe-rock interface shear 4.4. Post-analysis of modeling results with consideration for arching effect
stresses, i.e. ΣFy ≈ Fτ,cal. This implied that all the loads generated due to
the prescribed displacements were borne by the pipe-rock interface The use of a displaced ‘wish-in-place' pipe in the numerical analysis
elements and that there was no end bearing resistance incurred at the has been reliably demonstrated. The effects of arching on pipe-jacking
pipe ends, which was similarly observed in other simulations of pipe- forces remain to be addressed. From the earlier companion paper (Ong
jacking drives through highly weathered and fractured lithologies (Ong and Choo, 2018), it was found that the accrual of pipe-jacking forces
and Choo, 2016). This further confirmed the successful release of model through highly fractured rock masses was dependent on arching. The
boundary fixities at the pipe ends. The equivalence between ΣFy and excavation of a tunnel would instigate the occurrence of arching in the
Fτ,cal showed that equilibrium had been achieved between the pre-
scribed displacements, the incurred reaction loads and the rock-pipe
interface shear stresses, demonstrating compliance to the lower bound
theorem. This meant that prescribed displacements could be used to
obtain a safe and valid solution for the assessment of pipe-jacking
forces.
Fig. 10. Validation of numerical analysis: Case 3 – Uniform distribution of rock-
pipe interface shear stresses, τ1 (5 kN/m2) along longitudinal cross-section of
modeled pipeline.

7
C.S. Choo and D.E.L. Ong Engineering Geology 265 (2020) 105405

an arching ratio of less than 1.0 would imply the occurrence of active
arching, while an arching ratio of greater than 1.0 would indicate the
presence of passive arching. This fundamental understanding led to the
development of arching ratios primarily for studies conducted on em-
bankment fills (Borges and Marques, 2011; Deb, 2010; Han and Gabr,
2002; Ye et al., 2012).
For buried structures, Lee et al. (2006) expressed arching ratio with
the following equation
Δσv
AR = × 100%
σvo (9)
where AR is the arching ratio; Δσv is the change in vertical stress due to
tunneling; and σvo is the initial overburden pressure. From this, AR
could be applied to pipe-jacking by considering the normal stresses
acting on the pipeline. Thus, the arching ratio was proposed to be
γDe
σEV + 2
AR = × 100%
(
γ h+
De
2 ) (10)
This arching ratio, AR was based on an average point, at the depth
of the springline, denoted by the component γDe . By applying the pro-
2
posed arching ratio, AR (Eq. (8)) to the rock-pipe interface shear
stresses, the jacking force determined from τ1 is given as
JFτ , cal = AR × τ1 πDe (11)
The jacking forces determined from τ1 can be further verified
against the jacking forces determined from the reaction loads, ΣFy,
which is given as

Fig. 11. Validation of numerical analysis: Case 3 – Radial distribution of rock- ∑ Fy


JFFy, cal = AR ×
pipe interface shear stresses, τ1 (5 kN/m2) at mid-span of modeled pipeline. L (12)
In order to assess the accuracy of the numerical modeling technique,
upper soil or ‘soft rock’ mass, causing a redistribution of in-situ soil the calculated jacking forces, JFτ,cal and JFFy,cal were compared against
stresses and resulting in the reduction of stresses acting on the pipeline. the field measurements of jacking forces, JFmeas.
The use of a ‘wish-in-place' pipeline in the numerical model implied
that there was no overcut region separating the rock mass from the 5. Results of numerical analysis
pipeline. Therefore, it was proposed to apply an arching ratio manually
for the consideration of arching effects on the results from numerical Table 6 shows the results of the numerical modeling of Drives A, B
analysis. and C from the earlier companion paper (Ong and Choo, 2018), with
Arching ratios have been well-used to quantify the effects of the variations of these results against the measured jacking forces. In
arching. Based on the classical trapdoor experiment by Terzaghi (1936), the modeling of the studied Drives A, B and C, prescribed displacements
McNulty (1965) expressed arching ratio as the ratio of load on the were used to drive the modeled pipelines. For Drive A, the back-ana-
yielded trapdoor to that when the trapdoor is undeflected. The use of lyzed σEV was 12.5 kN/m2. Using Eq. (8), the resulting arching ratio, AR
such an arching ratio can be useful in quantifying states of arching, i.e. was 8.00%. This meant that the strength of the rock mass contributed to

Table 6
Analysis of finite element modeling results.
Drive A B C

Geology Shale Metagraywacke – siltstone Graywacke – phyllite


Prescribed displacement (mm) 100 100 100
Vertical stress acting on pipe crown, σEV (kN/m2) 12.5 −19.9a −38.5a
Arching ratio, AR (%)b 8.00 3.75 2.90
Measured jacking forces, JFmeas (kN/m) 37.2 20.7 23.6
Interface shear stress, τ1 (kN/m2) 65 104 159
Jacking forces calculated from interface shear stresses, JFτ,cal (kN/m)c 29.0 17.5 20.8
JFτ , cal − JFmeas −22.2 −15.4 −12.2
Variation, × 100 (%)
JFmeas
Reaction force, ΣFy (kN) 7233 8759 12,850
Jacking forces calculated from reaction force, JFFy,cal (kN/m)d 28.9 16.4 18.7
JFFy, cal − JFmeas −22.3 −20.8 −21.1
Variation, × 100 (%)
JFmeas
Difference in variations, 0.1 5.4 8.9
JFτ , cal − JFFy, cal
× 100 (%)
JFmeas

Note: a. Negative value of σEV was readjusted to zero.


b. From Eq. (8).
c. From Eq. (11).
d. From Eq. (12).

8
C.S. Choo and D.E.L. Ong Engineering Geology 265 (2020) 105405

a 92% reduction in normal loads acting on the pipe from the sur- analysis and the measured jacking forces correlated reasonably well,
rounding soil. From numerical analysis, the interface shear stress, τ1 thus demonstrating the reliability of the interpreted strength para-
was 65 kN/m2. Using Eq. (11), the jacking forces computed from τ1, meters, c't,p and ϕ′t,p through numerical modeling, as well that of the
JFτ,cal was 29.0 kN/m. Comparing this to the measured jacking force, back-analyzed parameters, μavg and σEV, through the consideration of
JFmeas of 37.2 kN/m, the calculated JFτ,cal underestimated JFmeas by arching effect.
22.2%. The reaction force, ΣFy measured from numerical analysis was
7233 kN. Using Eq. (12), the jacking forces calculated from ΣFy, JFFy,cal
was 28.9 kN/m, which underestimated the measured jacking forces by 6. Conclusions
22.3%. Thus, the results from numerical analysis of Drive A agreed
reasonably well with the measured jacking forces. From the earlier companion paper (Ong and Choo, 2018), three
For Drives B and C, the back-analyzed σEV values were negative. For pipe-jacking drives traversing highly fractured lithologies were studied.
the determination of arching ratios, AR, the negative σEV were adjusted The drives negotiated lithological units of shale (Drive A) and inter-
to be equal to zero (Choo and Ong, 2015; Terzaghi, 1943). The AR bedded lithological units of metagraywacke – siltstone (Drive B) and
values for Drives B and C were 3.75% and 2.90%, respectively. In greywacke – phyllite (Drive C) origins. From each of the studied pipe-
comparison with the arching ratio for Drive A, the AR values for Drives jacking drives, tunneling rock spoils were collected, reconstituted and
B and C were smaller, indicating that arching was more significant in subjected to direct shear testing. Peak tangential strength parameters
the latter drives. The heightened arching contributed to the measured c't,p and ϕ't,p were derived from the direct shear test results and used for
jacking forces, which were smaller in Drives B and C as compared to the backanalysis of the measured jacking forces, from which μavg and
those measured from Drive A (see Table 6). σEV were developed.
From the numerical modeling of Drive B, the pipe-rock interface Numerical modeling of the pipe-jacking drives was performed to
shear stresses, τ1 were 104 kN/m2, resulting in a calculated jacking ascertain the applicability of the parameters, c't,p, ϕ't,p, μavg and σEV re-
force, JFτ,cal of 17.5 kN/m. This was a 15.4% underestimation of the sulting from the earlier backanalysis of jacking forces. The developed
measured jacking forces of 20.7 kN/m. The reaction load, ΣFy was strength parameters, c't,p and ϕ't,p were applied to the modeled rock
8753 kN. The resulting calculated jacking force, JFFy,cal was 16.4 kN/m, mass surrounding the modeled pipelines. The concrete pipelines were
which was an underestimation of field measured jacking forces of modeled as ‘wish-in-place', with an increased material modulus. It was
20.8%. For Drive C, the pipe-rock interface shear stresses, τ1 were necessary to model the pipelines as rigid bodies in order to segregate
159 kN/m2. The jacking forces, JFτ,cal calculated from these interface the influence of axial loads from the pipe-rock interface shear stresses.
shear stresses was 20.8 kN/m, which was a 12.2% underestimation of The pipe-rock interface elements were modeled as cohensionless fric-
the measured jacking forces. From the reaction force, ΣFy of 12,850 kN, tional Mohr-Coulomb materials, with the strength parameters derived
the calculated jacking force, JFFy,cal was 18.7 kN/m, which under- from the backanalyzed μavg in Ong and Choo (2018). A low material
estimated the measured jacking forces by 21.1%. stiffness was used for the pipe-rock interface elements to facilitate the
It was useful to compare the results obtained from interface shear longitudinal displacement of the rigid pipeline. For the consideration of
stresses, τ1 and those obtained from reaction loads, JFFy,cal. The results arching effects in subsequent post-analysis, an arching ratio was de-
from τ1 were obtained from the mid-span of the modeled pipes, while veloped from σEV.
JFFy,cal was a measure of all the pipe-rock interface shear loads accrued The modeling technique was validated in accordance with the lower
along the pipeline, including end effects. Therefore, difference between bound theorem. When applying either prescribed displacements or
the results obtained from interface shear stresses, τ1 and those obtained loads for the propulsion of the modeled pipeline, the forces derived
from reaction loads, JFFy,cal could be used to assess the influence of end from the integration of interface shear stresses, τ1 were in static equi-
effects. This difference is given by. librium with the reaction loads, ΣFy. Thus, the stress field in the model
JFτ , cal − JFFy, cal was considered to be statically admissible, allowing for the modeling
Difference in variations = × 100% technique to be used for subsequent simulation of the studied pipe-
JFmeas (13)
jacking drives.
The differences in variations were generally small, with the highest The jacking forces derived from numerical analysis underestimated
difference exhibited in Drive C at 8.9%. This implied that end effects the measured jacking forces within a range of 15% to 22%. This was
had minimal influence on the accuracy of the numerical analysis. due to the use of the lower bound theorem for the assessment of the
Generally, the resulting τ1 and ΣFy from numerical analysis under- numerical modeling technique. Furthermore, the cementation of the
estimated field measurements of jacking forces within a range of 15% to tunneling rock spoils was reduced through crushing of the rock mass.
22%. This was attributed to the use of the lower bound approach The idealization of an average pipeline traversing homogeneous geo-
adopted in the current study for the numerical modeling of the studied logical conditions contributed to the underestimation of jacking forces.
pipe-jacking drives. Furthermore, the peak tangential strength para- However, the jacking forces obtained from τ1 and ΣFy in the numerical
meters, c't,p and ϕ't,p were developed from direct shear testing of re- analysis were considered to be in reasonable agreement with the
constituted tunneling rock spoils in order to simulate the highly frac- measured jacking forces.
tured and highly weathered nature of the in-situ ‘soft rock’, as described The findings showed that reliable strength parameters can be de-
in the earlier companion paper (Ong and Choo, 2018). These rock spoils veloped from the direct shear testing of reconstituted tunneling rock
were crushed from the traversed lithologies during pipe-jacking, thus spoils to be subsequently used for the assessment of jacking forces for
diminishing the natural cementation of the rock mass and resulting in pipe-jacking drives negotiating highly weathered and fractured ‘soft
lower bound solutions (Ong and Choo, 2016). Additionally, the as- rock’ such as those encountered in the Tuang Formation in Kuching,
sessment of pipe-jacking forces depended on the use of the Pellet- Malaysia. This reliability has been demonstrated through numerical
Beaucour and Kastner (2002) jacking force model, which required the modeling of the pipe-jacking drives using results obtained from direct
assumption of average uniform conditions along the modeled pipelines. shear tests from the earlier companion paper Ong and Choo (2018).
This assumption was applied to the strength and stiffness properties of This method of back-analysis could allow for the prediction of jacking
the modeled rock mass, the pipeline and the pipe-rock interface ele- forces in highly weathered, highly fractured lithologies for future pipe-
ments. This idealization consequently required the use of the ‘wish-in- jacking drives, through the use of direct shear tests on reconstituted
place’ technique for introducing the pipeline into the model, which crushed rock.
contributed to the underestimation of jacking forces. Despite these
underestimations, the jacking forces computed from the numerical

9
C.S. Choo and D.E.L. Ong Engineering Geology 265 (2020) 105405

Declaration of Competing Interest numerical analysis of cutting rock and soil by TBM cutting tools. Tunn. Undergr.
Space Technol. 81, 428–437. https://doi.org/10.1016/j.tust.2018.08.015.
Li, C., Zhong, Z., Liu, X., Tu, Y., He, G., 2019. Numerical simulation for an estimation of
The authors declare that they have no known competing financial the jacking force of ultra-long-distance pipe jacking with frictional property testing at
interests or personal relationships that could have appeared to influ- the rock mass–pipe interface. Tunn. Undergr. Space Technol. 89, 205–221. https://
ence the work reported in this paper. doi.org/10.1016/j.tust.2019.04.004.
McNulty, J.W., 1965. An Experimental Study of Arching on Sand, in US Army Corps of
Engineers Technical Report I-674. U.S. Waterways Experimental Station, Vicksburg.
Acknowledgment Merifield, R.S., Sloan, S.W., 2006. The ultimate pullout capacity of anchors in frictional
soils. Can. Geotech. J. 43 (8), 852–868. https://doi.org/10.1139/t06-052.
Ong, D.E.L., Choo, C.S., 2016. Back-analysis and finite element modeling of jacking forces
The Authors would like to express their thanks for the financial in weathered rocks. Tunn. Undergr. Space Technol. 51, 1–10. https://doi.org/10.
generosity shown by Hock Seng Lee Bhd. and Jurutera Jasa (Sarawak) 1016/j.tust.2015.10.014.
Sdn. Bhd. during this study. Ong, D.E.L., Choo, C.S., 2018. Assessment of non-linear rock strength parameters for the
estimation of pipe-jacking forces. Part 1. Direct shear testing and backanalysis. Eng.
Geol. https://doi.org/10.1016/j.enggeo.2018.07.013.
References Pellet-Beaucour, A.L., Kastner, R., 2002. Experimental and analytical study of friction
forces during microtunneling operations. Tunn. Undergr. Space Technol. 17 (1),
Augarde, C.E., Lyamin, A.V., Sloan, S.W., 2003. Prediction of undrained sinkhole col- 83–97. https://doi.org/10.1016/S0886-7798(01)00044-X.
lapse. J. Geotech. Geoenviron. 129 (3), 197–205. https://doi.org/10.1061/(ASCE) Plaxis bv, 2013. PLAXIS 3D 2013 Manual. (Delft, The Netherlands).
1090-0241(2003)129:3(197). Potts, D.M., Zdravković, L., 1999. Finite Element Analysis in Geotechnical Engineering:
Barla, M., Camusso, M., Aiassa, S., 2006. Analysis of jacking forces during microtunnel- Theory. Vol. 1 Thomas Telford Services Limited.
ling in limestone. Tunn. Undergr. Space Technol. 21 (6), 668–683. https://doi.org/ Pusch, R., Börgesson, L., 1998. Strain parameters of discontinuities in rock for finite
10.1016/j.tust.2006.01.002. element calculation. Eng. Geol. 49 (3), 201–206. https://doi.org/10.1016/S0013-
Barton, N., Lien, R., Lunde, J., 1974. Engineering classification of rock masses for the 7952(97)00050-1.
design of tunnel support. Rock Mech. 6 (4), 189–236. https://doi.org/10.1007/ Randolph, M.F., Houlsby, G.T., 1984. The limiting pressure on a circular pile loaded
BF01239496. laterally in cohesive soil. Géotechnique 34, 613–623. https://doi.org/10.1680/geot.
Basarir, H., Ozsan, A., Karakus, M., 2005. Analysis of support requirements for a shallow 1984.34.4.613.
diversion tunnel at Guledar dam site, Turkey. Eng. Geol. 81 (2), 131–145. https:// Shou, K.J., Yen, J., Liu, M., 2010. On the frictional property of lubricants and its impact
doi.org/10.1016/j.enggeo.2005.07.010. on jacking force and soil–pipe interaction of pipe-jacking. Tunn. Undergr. Space
Bieniawski, Z.T., 1979. Tunnel design by rock mass classifications, in US Army Corps of Technol. 25 (4), 469–477. https://doi.org/10.1016/j.tust.2010.02.009.
Engineers Technical Report GL-79-19. U.S. Waterways Experimental Station, Smith, C.C., 1998. Limit loads for an anchor/trapdoor embedded in an associative
Vicksburg. Coulomb soil. Int. J. Numer. Anal. Methods Geomech. 22 (11), 855–865. https://doi.
Bieniawski, Z.T., 1989. Engineering Rock Mass Classifications: a Complete Manual for org/10.1002/(SICI)1096-9853(199811)22:11%3C855::AID-NAG945%3E3.0.CO;2-4.
Engineers and Geologists in Mining, Civil, and Petroleum Engineering. Wiley. Smith, C.C., 2012. Limit loads for a shallow anchor/trapdoor embedded in a non-asso-
Borges, J.L., Marques, D.O., 2011. Geosynthetic-reinforced and jet grout column-sup- ciative Coulomb Soil. Géotechnique 62 (7), 563–571. https://doi.org/10.1680/geot.
ported embankments on soft soils: Numerical analysis and parametric study. Comput. 10.P.136.
Geotech. 38 (7), 883–896. https://doi.org/10.1016/j.compgeo.2011.06.003. Terzaghi, K., 1936. Stress distribution in dry and in saturated sand above a yielding trap-
Choo, C.S., Ong, D.E.L., 2015. Evaluation of pipe-jacking forces based on direct shear door. In: Proc., 1st Intl. Conf. on Soil Mechanics and Foundation Engineering.
testing of reconstituted tunneling rock spoils. J. Geotech. Geoenviron. Eng. 141 (10). Graduate School of Engineering, Harvard Univ., Cambridge, Massachusetts, USA, pp.
https://doi.org/10.1061/(ASCE)GT.1943-5606.0001348. 307–311.
Davis, E.H., Gunn, M.J., Mair, R.J., Seneviratne, H.N., 1980. The stability of shallow Terzaghi, K., 1943. Theoretical Soil Mechanics. John Wiley & Sons, Inc.
tunnels and underground openings in cohesive material. Géotechnique 30 (4), Tian, Y., Cassidy, M.J., Randolph, M.F., Wang, D., Gaudin, C., 2014. A simple im-
397–416. https://doi.org/10.1680/geot.1980.30.4.397. plementation of RITSS and its application in large deformation analysis. Comput.
Deb, K., 2010. A mathematical model to study the soil arching effect in stone column- Geotech. 56, 160–167. https://doi.org/10.1016/j.compgeo.2013.12.001.
supported embankment resting on soft foundation soil. Appl. Math. Model. 34 (12), Wang, X., Wang, L.B., 2006. Continuous interface elements subject to large shear de-
3871–3883. https://doi.org/10.1016/j.apm.2010.03.026. formations. Int. J. Geomech. 6 (2), 97–107. https://doi.org/10.1061/(ASCE)1532-
Ding, W., Li, S., 2011. Study on Damage of Jointed Rock Tunnel by Stochastic Fractures. 3641(2006)6:2(97).
Instrumentation, Testing, and Modeling of Soil and Rock Behavior 47–55. https:// Valliappan, S., Ang, K.K., 1988. Finite element analysis of vibrations induced by propa-
doi.org/10.1061/47633(412)7. gating waves generated by tunnel blasting. Rock Mech. Rock Eng. 21 (1), 53–78.
Dounias, G.T., Potts, D.M., 1993. Numerical analysis of drained direct and simple shear https://doi.org/10.1007/BF01019675.
tests. J. Geotech. Eng. 119 (12), 1870–1891. https://doi.org/10.1061/(ASCE)0733- Wang, D., Hu, Y., Randolph, M.F., 2009. Three-dimensional large deformation finite-
9410(1993)119:12(1870). element analysis of plate anchors in uniform clay. J. Geotech. Geoenviron. 136 (2),
Finno, R.J., Lawrence, S.A., Allawh, N.F., Harahap, I.S., 1991. Analysis of performance of 355–365. https://doi.org/10.1061/(ASCE)GT.1943-5606.0000210.
pile groups adjacent to deep excavation. J. Geotech. Eng. 117 (6), 934–955. https:// Ye, G., Zhang, Z., Han, J., Xing, H., Huang, M., Xiang, P., 2012. Performance evaluation of
doi.org/10.1061/(ASCE)0733-9410(1991)117:6(934). an embankment on soft soil improved by deep mixed columns and prefabricated
Genis, M., Basarir, H., Ozarslan, A., Bilir, E., Balaban, E., 2007. Engineering geological vertical drains. J. Perform. Constr. Facil. 27 (5), 614–623. https://doi.org/10.1061/
appraisal of the rock masses and preliminary support design, Dorukhan Tunnel, (ASCE)CF.1943-5509.0000369.
Zonguldak, Turkey. Eng. Geol. 92 (1), 14–26. https://doi.org/10.1016/j.enggeo. Yen, J.H., Shou, K.J., 2015. Numerical simulation for the estimation the jacking force of
2007.02.005. pipe jacking. Tunn. Undergr. Space Technol. 49, 218–229. https://doi.org/10.1016/
Georgiadis, K., Sloan, S.W., Lyamin, A.V., 2013. Undrained limiting lateral soil pressure j.tust.2015.04.018.
on a row of piles. Comput. Geotech. 54, 175–184. https://doi.org/10.1016/j. Yin, Z.-Z., Zhu, H., Xu, G.-H., 1995. A study of deformation in the interface between soil
compgeo.2013.07.003. and concrete. Comput. Geotech. 17 (1), 75–92. https://doi.org/10.1016/0266-352X
Han, J., Gabr, M., 2002. Numerical analysis of geosynthetic-reinforced and pile-supported (95)91303-L.
earth platforms over soft soil. J. Geotech. Geoenviron. 128 (1), 44–53. https://doi. Zhang, X., Krabbenhoft, K., Pedroso, D.M., Lyamin, A.V., Sheng, D., da Silva, M.V., Wang,
org/10.1061/(ASCE)1090-0241(2002)128:1(44). D., 2013. Particle finite element analysis of large deformation and granular flow
Hu, L., Pu, J., 2004. Testing and modeling of soil-structure interface. J. Geotech. problems. Comput. Geotech. 54, 133–142. https://doi.org/10.1016/j.compgeo.2013.
Geoenviron. 130 (8), 851–860. https://doi.org/10.1061/(ASCE)1090-0241(2004) 07.001.
130:8(851). Zhang, L.Q., Yue, Z.Q., Yang, Z.F., Qi, J.X., Liu, F.C., et al., 2006. A displacement-based
Itasca Consulting Group Inc, 1999. UDEC Universal Distinct Element Code, version 3.10. back-analysis method for rock mass modulus and horizontal in situ stress in tunneling
User's Guide. (Minneapolis). – Illustrated with a case study. Tunn. Undergr. Space Technol. 21 (6), 636–649.
Itasca Consulting Group Inc, 2001. FLAC Fast Lagrangian Analysis of Continua, Version https://doi.org/10.1016/j.tust.2005.12.001.
4.0. User’s Guide. (Minneapolis). Zhu, W., Li, S., Li, S., Chen, W., Lee, C.F., 2003. Systematic numerical simulation of rock
Lee, C.J., Wu, B.R., Chen, H.T., Chiang, K.H., 2006. Tunnel stability and arching effects tunnel stability considering different rock conditions and construction effects. Tunn.
during tunneling in soft clayey soil. Tunn. Undergr. Space Technol. 21 (2), 119–132. Undergr. Space Technol. 18 (5), 531–536. https://doi.org/10.1016/S0886-7798(03)
https://doi.org/10.1016/j.tust.2005.06.003. 00070-1.
Li, G., Wang, W., Jing, Z., Zuo, L., Wang, F., Wei, Z., et al., 2018. Mechanism and

10

You might also like