Effect of interparticle behavior on the development of soil arching in soil structure interaction

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Tunnelling and Underground Space Technology 106 (2020) 103610

Contents lists available at ScienceDirect

Tunnelling and Underground Space Technology


journal homepage: www.elsevier.com/locate/tust

Effect of interparticle behavior on the development of soil arching in soil- T


structure interaction
M.I. Peeruna, D.E.L. Ongb, , C.S. Chooc, W.C. Chengd

a
Griffith University, School of Engineering and Built Environment, 170 Kessels Road, Nathan, Queensland 4111, Australia
b
Griffith University, School of Engineering and Built Environment, 170 Kessels Road, Nathan, Queensland 4111, Australia
c
Research Centre for Sustainable Technologies, Faculty of Engineering, Science & Computing, Swinburne University of Technology Sarawak Campus, 93350 Kuching,
Sarawak, Malaysia
d
Shaanxi Key Laboratory of Geotechnical and Underground Space Engineering, School of Civil Engineering, Xi'an University of Architecture and Technology, China

ARTICLE INFO ABSTRACT

Keywords: Direct shear tests have been commonly used to study the frictional stresses during pipe-jacking. Particle angu­
Apparent cohesion larity and mineralogy will influence the shear strength with apparent cohesion contributing to the arching effect
Direct shear during soil-structure interaction via pipe-jacking. Past researchers found it challenging to physically study
GeoPIV particle behavior along the shear band during shearing due to the opaque nature of the equipment. This study
Interlocking
used an original transparent shear box to conduct direct shear tests on tunneling rock spoils of sandstone and
Soil arching
shale. Sequential images were captured during shearing and analyzed using GeoPIV software to demonstrate
Pipe-jacking
localized activities which were found to influence the apparent cohesion. Furthermore, the findings were suc­
cessfully used to assess the arching phenomenon observed during pipe-jacking and other established pipe-
jacking case studies. Rounded particles with strong mineralogy were found to be more likely to produce an
arching effect as compared to angular particles with weaker minerals.

1. Introduction which reduces the vertical stresses at the pipe crown. The reduction of
vertical load on the jacking pipe due to arching, would strongly influ­
Pipe-jacking consists of building an underground pipeline using ence the jacking forces. Hence, it is essential to define the vertical soil
segmental pipes. These segmental pipes are lowered through a stresses acting on the jacking pipes for the estimation of jacking forces
launching shaft, behind a micro-tunnel boring machine (MTBM) which during the design process of the project (Misra et al., 2008; Rahjoo
is used to bore through the ground. Hydraulic jacks are used to thrust et al., 2012; Olson, 2013; Olson et al., 2016; Khondoker et al., 2016).
the segmental pipes and the microtunnel boring machine through the Several jacking force models are currently available and used to esti­
ground (Shen et al. 2016). During pipe-jacking works, jacking forces mate jacking forces. Among the various jacking force models, some
have been found to be dependent on the traversed geological strata have been acquired statistically by Chapman and Ichioka (1999), em­
(Cheng et al., 2018b). Better understanding of jacking forces, including pirically by Osumi (2000) and experimentally by Staheli (2006) as
advancement in soil testing (Mehdizadeh et al. 2016, 2017) and mod­ described by Choo and Ong (2015). However, those models are parti­
eling techniques (Ong et al. 2006) in soil-structure interaction problems cularly dependent on the traversed soil types, with minimal con­
for specific geological strata (Ong et al. 2003, Ong and Choo 2011, siderations for pipe-jacking through highly weathered rock masses. The
2018) would lead to a more economical and sustainable project. jacking force models are limited in terms of ability to characterize the
Arching phenomenon is of great interest during tunneling where the pipe-jacking conditions (W.-C. Cheng et al., 2018; W. Cheng et al.,
lithology over the tunnel could arch, thus potentially allowing the 2018). Choo and Ong (2015) pointed out an established jacking force
tunnel to self-stand without fully collapsing onto the jacking pipes. model developed by Pellet-Beaucour and Kastner (2002) which con­
Such tunnels with favorable arching can self-stand due to redistribution siders arching. To cater for this effect, cohesion, C’ and internal friction
of stresses around the tunnel and pipe surfaces. Arching is first de­ angle, ϕ’ of the surrounding soils are required.
scribed as a redistribution of geostatic stresses by Terzaghi (1943) Several researchers have studied jacking forces by conducting direct

Corresponding author at: Griffith University, School of Engineering and Built Environment, 170 Kessels Road, Nathan, Queensland 4111, Australia.

E-mail addresses: irfaan.peerun@griffithuni.edu.au (M.I. Peerun), d.ong@griffith.edu.au (D.E.L. Ong), cschoo@swinburne.edu.my (C.S. Choo),
w-c.cheng@xauat.edu.cn (W.C. Cheng).

https://doi.org/10.1016/j.tust.2020.103610
Received 8 March 2020; Received in revised form 11 August 2020; Accepted 28 August 2020
0886-7798/ © 2020 Elsevier Ltd. All rights reserved.
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

shear tests for the evaluation of soil-structure frictional stresses. Staheli are: 1) to assess the influence of particle angularity and mineralogy
(2006) performed interface direct shear test using sand and various contributing to apparent cohesion, 2) to evaluate the vertical load
pipe materials with different surface roughnesses. Based on the find­ acting on the jacking pipe based on the shear strength and particle
ings, a frictional jacking force model was developed but it was limited behavior which is indicative of arching effect, and 3) to estimate
to unlubricated drives only. However, direct shear test was able to jacking forces using the shear strength parameters and compare against
produce the strength parameters, which were used for the assessment of jacking force records from the construction site.
jacking forces. Shou et al. (2010) studied the frictional properties of
several lubricants by simulating pipe-soil interaction during pipe- 2. Assessment of jacking forces
jacking by conducting a simple large-scale frictional test. By adopting
the frictional coefficient obtained from the experiment, the assessment Several jacking force models are currently available for the pre­
of jacking forces showed an overestimation when compared to site re­ diction of jacking forces in pipe-jacking works. The jacking force
cords, which was due to overestimating the pipe-soil contact area. equations were derived statistically (Chapman and Ichioka, 1999),
Nonetheless, the experiment showed substantial results for the assess­ empirically (Osumi, 2000), and experimentally (Staheli, 2006) which
ment of jacking forces using lubrication. Ong and Choo (2016), Choo were used for drives passing through sand or clay. Hence, there were
and Ong (2020) and Ong and Choo (2018) conducted direct shear test limited considerations for drives traversing rocks. The jacking force
and finite element modeling for the assessment of jacking forces model developed by Pellet-Beaucour and Kastner (2002) took into
through drives in highly fractured and weathered rock. The strength consideration the soil arching phenomenon. Soil arching was first stu­
properties of the reconstituted rock spoils were incorporated into the died through a trap door experiment by Terzaghi (1936) and was de­
jacking force model (Pellet-Beaucour and Kastner, 2002) to produce scribed as the relaxation of soil stresses due to disturbance in geostatic
reliable jacking force analyzes. Hence, by reconstituting the tunneling stresses.
rock spoils, direct shear test results can be used to assess the jacking Sterling (2020) described that several jacking force models in­
forces for different geologic lithologies. However, to have a better in­ corporated the findings from Terzaghi’s trap door experiment
sight on the arching effect, better understanding is required at the (Terzaghi, 1943) for the assessment of soil pressure acting on the pipe.
particulate level during shearing. It is important to demonstrate the The geometry and soil conditions would influence the shear band for­
non-continuous and non-uniform material during the arching me­ mation during pipe-jacking cases which may not fully match with the
chanism of soil, at the particle size scale (Zhu and Gong, 2014). trap door experiment. Arching depends on critical factors namely the
DeJong and Westgate (2009) conducted an interface direct shear silo width, height of shearing bands, friction angle and cohesion (Zhang
test to study the localized mechanism during load transfer from a et al. 2016). However, not all jacking force models take into con­
structure to the surrounding soil. Particle image velocimetry (PIV) was sideration the vertical load acting onto the pipe. Soil arching was stu­
used to quantify the displacement and strains of the soil particles during died by Choo and Ong (2015) and Ong and Choo (2016) for pipe-
the soil-structure interface shearing. The main factors influencing the jacking drives through highly weathered ‘soft rock’ masses by adopting
local interface behavior were the relative density of the soil, the soils the Pellet-Beaucour and Kastner (2002) jacking force model. The au­
particle angularity and hardness, and the surface roughness of the thors contextualised ‘soft rock’ as highly weathered rock masses which
structure. Fukuoka et al. (2006) conducted a ring shear test using a exhibited soil-like behaviour, i.e. arching. The approach took into ac­
transparent shear box to study the shear zone formation in granular count the vertical rock stresses (σEV) acting vertically on the pipe crown,
materials with the aid of particle image velocimetry (PIV) technology. which was used to assess the degree of arching. A negative value of σEV
Shear zone thickness was studied for drained and undrained coarse- would indicate the occurrence of arching favorable for lower pipe-
grained silica sands. The test specimens were sheared at a speed of jacking forces, while a positive value of σEV would indicate unfavorable
10 cm/s and 100 cm/s. However, a complete analysis of the whole arching, resulting in larger frictional jacking forces. By incorporating
shear zone was not achieved due to partial obstruction at the shear gap, the Mohr-Coulomb parameters, i.e. apparent cohesion and friction
which was made of a non-transparent material. Velocity distribution angle, obtained from the direct shear test results into the jacking force
profiles during shearing was produced from the ring shear tests and the model, an indication of arching was made and subsequently validated
data for the invisible section along the shear band was interpolated. No against site measurements of jacking forces, lubrication, jacking speed,
difference was found in shear zone thickness between the drained and and other significant pipe-jacking activities. For example, case studies
undrained tests. from Choo and Ong (2020) recorded a vertical stress of 12.5 kN/m2
There have been ongoing efforts in providing better insight on acting on the pipe when tunneling through a shale drive. A meta­
particle behavior during shearing. Due to the opaque nature of the graywacke-siltstone drive and graywacke-phyllite drive produced
shear equipment, it is challenging to physically observe the inter­ σEV = -19.9 kN/m2 and σEV = -38.5 kN/m2, respectively. This de­
particle activities during shearing. To overcome this, numerical mod­ monstrates that there was no vertical load acting on the pipe for the two
elling has been vastly adopted for the study of discrete particles (Jing drives resulting in significant arching as opposed to the least favorable
et al., 2017; Wang et al., 2019; Zhou et al., 2019; Zhu et al., 2019). arching condition for the shale drive. Case studies from Ong and Choo
Particle angularity and particle mineralogy were found to be the major (2016) also demonstrated the unfavorable arching condition for shale
influencing factors of arching. Hence, although numerical models have drive with a vertical stress, σEV = 11.7 kN/m2. Other drives in sand­
made significant contribution, there is still a need to physically study stone and phyllite were more favorable to arching with σEV = −20.8
particles activities such as interlocking and breakages which would kN/m2 and σEV = –22.3 kN/m2 for the respective drives.
influence apparent cohesion and subsequently arching. By further un­ Table 1 shows the suitability of jacking force models with con­
derstanding the influence of arching on pipe-jacking forces for various sideration for the vertical soil stress acting on the pipe as an indication
geologic lithologies, contribution to a more economical and efficient of arching. The jacking force equations along with its parameters have
pipe-jacking design can be achieved. An efficient assessment of jacking been listed for the models developed by Chapman and Ichioka (1999),
force would allow longer jacking distance and avoid damages to sur­ Osumi (2000), Staheli (2006) and Pellet-Beaucour and Kastner (2002).
rounding structures due to excessive jacking force (Cheng et al., 2019). From the various jacking force equations, only Pellet-Beaucour and
This paper discusses the effects of particle angularity and miner­ Kastner (2002) jacking force model took into consideration of the
alogy on the strength of tunneling spoils of sandstone and shale during vertical stress acting on the pipe. The said model is also applicable to a
shearing. This was achieved by using a unique purpose-built trans­ wider range of ground condition including highly weathered ‘soft
parent shear box to capture sequential images of the particles, which rocks’. Hence, the jacking force model developed by Pellet-Beaucour
were then analyzed using PIV technology. The objectives of this study and Kastner (2002) was used in this study to assess jacking forces for

2
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

Table 1
Suitability of jacking force models for the assessment of soil arching.
Jacking force model Chapman and Osumi (2000) Staheli (2006) Pellet-Beaucour and Kastner (2002)
Ichioka (1999)

Equation P = a + 0.38D fo = ( Bc q + w ) µ + Bc C
JFfrict = µint ·
·r·cos 45 + r
2
· ·d·l
F = µLDe
2 ( EV +
De
2 )+K (
2 EV +
De
2 )
tan r
Parameters P = frictional Fo = frictional component of μint = pipe-soil residual interface F = total frictional jacking force;
resistance along the jacking force; friction coefficient; L = pipe span;
pipe run; β = jacking force reduction γ = total unit weight of the soil; De = outer pipe diameter;
a = parameter factor; ϕr = residual friction angle of the γ = soil unit weight;
dependent on soil Bc = outer diameter of pipe; soil; h = soil cover from the ground level to the pipe
type; and q = normal force; d = pipe diameter; crown;
D = external μ’ = interface friction coefficient, r = pipe radius; and K = lateral earth pressure coefficient; and
diameter of pipe. ϕ = interface friction angle; l = pipe length. K2 = thrust coefficient of soil acting on the pipe,
w = pipe weight; and σEV is the vertical stress acting on the pipe crown
C’ = pipe-soil adhesion
Applicability to ground Clay, sand and Clay, sand and gravels Clean sands, non-lubricated drives Clay, sand, gravel and rock; lubricated drives
condition gravels
Consideration of Not applicable Not applicable Not applicable Applicable
vertical soil stress
acting onto pipe
crown (soil
arching)

drives passing through geologic lithologies namely sandstone and shale. Ong (2015), Ong and Choo (2018) and Choo and Ong (2020) have
The jacking force model is defined as successfully used the Pellet-Beaucour and Kastner (2002) model to as­
sess the vertical stress acting onto the pipe. The results were compared
De De
F = µLDe EV + + K2 EV + to site data and fair agreement was achieved.
2 2 2 (1) Frictional resistance is obtained by multiplying the total normal force
where acting on the pipes by an effective friction coefficient, µ. The coefficient of
friction will differ as a function of soil nature, outer pipe surface roughness
F = total frictional jacking force; and lubrication (Pellet-Beaucour and Kastner, 2002). Lubrication slurry
µ = coefficient of friction; consist mainly of bentonitic clay and their compositions are usually adjusted
L = pipe span; on site. The use of lubrication can significantly reduce frictional stresses
De = outer pipe diameter; during pipe-jacking up to 77% of the initial frictional stress. The efficiency
σEV = vertical soil stresses acting on the pipe crown; of lubrication is subjected to the volume of lubricant injected, the process of
γ = soil unit weight; injection, and the soil and lubricant nature. For jacking drives using lu­
K2 = thrust coefficient of soil acting onto the pipe (recommended brication, Stein (2005) suggested to adopt a coefficient of soil-pipe friction
value of 0.3 by Pellet-Beaucour and Kastner, 2002). ranging from 0.1 to 0.3.

The indicative vertical soil stresses, σEV, was obtained from Eq. (2). 3. Materials and methods

b ( 2C
) 3.1. Tested materials
EV =
2K tanØ
b
(1 e 2K h tan
b
) (2)
The test specimens consisted of spoils of sandstone and shale ob­
where tained from pipe-jacking sites in Kuching, Malaysia. Based on visual
inspection using a laboratory microscope, sandstone particles were
C = soil cohesion; categorized as rounded to sub-angular smooth particles, while shale
ϕ = soil internal friction angle; particles were defined as angular with rougher surface. Pfeffer (2014)
γ = soil unit weight; described shale as relatively weaker due to being fissile, known as ea­
K = lateral earth pressure coefficient (recommended value of 1.0 by sily split apart due to weathering. Petrographic analysis showed that
Pellet-Beaucour and Kastner, 2002); the shale spoils were mainly clay minerals, silt-sized quartz grains, tiny
h = soil cover from the ground level to the pipe crown flakes of mica and carbonaceous materials. Brownish iron oxide veins
and quartz veins occurred as infilling along the cleavage. The relatively
and b is the influencing soil width above the pipe, defined as strong particles of sandstone consisted high amount of pure quartz and
is categorized as orthoquartzite (Pfeffer, 2014).
b = De 1 + 2tan
4 2 (3)
3.2. Direct shear testing
The normal stress acting onto the jacked pipe is a combined product
from the surrounding soil and the pipe weight (Pellet-Beaucour and In order to obtain the strength parameters of sandstone and shale spoils,
Kastner, 2002). The tunnel boring machine would influence the initial direct shear tests were conducted according to ASTM D3080-11 (ASTM
state of stresses around the tunnel resulting in soil relaxation due to the 2011). The strength caracteristics, namely friction angle and apparent co­
creation of an overcut. Terzaghi’s silo model assumes that the ground is hesion, were essential for the assessment of jacking forces using the Pellet-
moving along two vertical planes from the top of the pipe and that the Beaucour & Kastner jacking force model (Pellet-Beaucour and Kastner,
movement is significant enough to create sliding planes. Based on this 2002). To have a better understanding on the behaviors of the particles
assumption, Pellet-Beaucour and Kastner (2002) produced Eq. (2) for along the shear band, the shearing process was explained through the use of
the assessment of vertical soil stress acting on the pipe crown. Choo and a four-stage shearing model which was developed by Li and Aydin (2010).

3
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

3.2.1. Shearing stages 3.3.2. GeoPIV


Li and Aydin (2010) described Stage 1 as ‘end zone deformation’ White and Take (2002) developed GeoPIV, which is a MATLAB
when shearing starts and the particles rearrange themselves within module that uses particle image velocimetry (PIV) technique for geo­
existing voids causing the sample to contract. Stage 1 ends at the lowest technical testing applications. From a batch of sequential images, the
point of contraction where particles interlock, resulting in the increase software uses the concept of PIV to produce displacement vectors to
of inter-particle contact points, thus preventing any further contraction indicate the movement and speed of particles when subjected to a
of the test specimen (Shimizu, 1997). Stage 2 known as ‘particle in­ variety of tests. Many researchers have used GeoPIV to study particles
terlocking’ starts from the lowest contraction point where particles in- movement in various fields (White and Take, 2002; DeJong and
contact with each other have to surpass interlocking along the shear Westgate, 2009; Jackson, 2010; Grognet, 2011; Kelly, 2014). This
band resulting in localized dilation. This stage ends at the peak shear software is an accurate tool which can be used to better understand the
stress where maximum interlocking has been achieved within the shear behaviors of the particles during shearing. Localized activities within
band. Stage 3 is the ’shear zone formation’ which starts from the peak the test specimen, such as localized dilation and compression, can be
shear stress where particles rearrange themselves into the voids by ei­ captured to understand the mechanisms that contribute to friction angle
ther rolling or rotating, producing a reduction in shear resistance. In and apparent cohesion of the tested material.
direct shear testing, this is usually observed as a reduction from peak
shear stresses towards residual shear stresses. The behaviors of the
3.4. Testing procedure
particles will produce a looser shear band layer (Oda and Konishi, 1974;
Fukuoka et al., 2006). This stage ends when a constant residual stress is
3.4.1. Sample preparation
achieved. ‘Steady shear’ is the last stage (Stage 4) of shearing which is
Tunneling spoils of sandstone and shale were oven-dried and
achieved by constant shear stress where minimal vertical deformation is
scalped to obtain a particle size distribution (PSD) passing 2.36 mm
recorded. The particles slide along the shear band in an equilibrium of
sieve and retained by 75 µm sieve. Scalping was required by ASTM
dilation and contraction while the bottom half of the shear box dis­
D3080-11 (ASTM, 2011), which stated that the PSD of the tested spe­
places with constant amplitude and attain residual state (Li and Aydin,
cimen was restricted to a minimum specimen width not less than ten
2010).
times the maximum particle size diameter, dmax and the minimum
specimen height not less than 6dmax. Choo and Ong (2015) have pre­
3.3. Experimental setup viously used the scalping method for shearing reconstituted tunneling
rock spoils.
A fully automated shearing equipment was used to conduct direct Sandstone spoils was categorized as well-graded while the spoils of
shear tests on the two specimens up to a maximum horizontal dis­ shale was categorized as poorly-graded spoils according to Unified Soil
placement of 15 mm. This was to ensure that the specimens had Classification System (ASTM, 2000). Dry pluviation technique was
reached a residual state, i.e. Stage 4 of shearing. The tests were con­ adopted to achieve dense test specimens using a funnel of 35 mm
ducted at normal stresses of 100 kPa, 300 kPa and 500 kPa to develop opening and 100 mm fall height (DeGregorio, 1990). To ensure re­
the respective failure criterion for each material. peatability, similar dry density was maintained for each specimen when
tested at various normal loads. Sandstone and shale spoils produced a
dry density of 1,995.4 kg/m3 and 1,819.4 kg/m3 respectively. Table 2
3.3.1. Purpose-built transparent shear box
shows the physical characteristics of the test specimens.
A purpose-built shear box made of transparent Perspex was used to
study the behaviors of the particles during shearing. The transparent
shear box allowed for clear visibility of the particles along the shear 3.4.2. Image acquisition
band (see Fig. 1), which overcomes the challenges from past studies Sequential images were captured using a Canon EOS 450D camera
(Wang and Sassa, 2002; Wafid et al., 2004; Fukuoka et al., 2006). The while the specimen was sheared for a total horizontal displacement of
size and specifications of the transparent shear box were made to 15 mm. The camera was remotely controlled to avoid any disturbance
conform with ASTM D3080-11 (ASTM, 2011). A bottom plate was fixed during the image acquisition. Dim diffused lights were used to improve
to the shear box to ensure that the latter was raised above the water visibility of the particles while preventing the casting of shadows onto
reservoir to obtain an unobstructed view during image acquisition. the shear box. The equipment setup is shown in Fig. 2.

Fig. 1. Sample image with region of interest highlighted obtained from transparent shear box.

4
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

Table 2
Physical characteristics of test specimens.
Specimen D60 (mm) D30 (mm) D10 (mm) Coefficient of uniformity Coefficient of curvature Dry density (Kg/ Unified Soil Classification System (ASTM
Cu Cc m3) 2000)

Sandstone spoils 0.70 0.27 0.10 7.00 1.04 1995.4 Well-graded sand-sized spoils
Shale spoils 0.90 0.40 0.20 4.50 0.89 1819.4 Poorly graded sand-sized spoils

Fig. 2. Equipment setup for image acquisition.

3.4.3. Region of interest dilatancy angle for sandstone and shale. Eq. (4) was used to obtain the
Prior to the GeoPIV analysis, grids were drawn onto the images dilatancy angle, , as defined by Bolton (1986).
known as patches to produce a mesh file. Higher image resolution
would produce larger number of patches and hence more information d y
tan =
can be acquired (Grognet, 2011). The region of interest (ROI) and d yz (4)
image dimensions of 4272 × 2848 pixels obtained from the 72-dpi
resolution camera are shown in Fig. 1. The ROI was set constant along where y is the vertical strain and yz is the shear strain. Both vertical
the shear gap for all the tests to obtain a constant image of the upper strain and shear strain were recorded at the occurrence of peak shear
and lower halves of the shear box. stress during direct shear tests. A positive dilatancy angle would mean
that the specimen is dilating while a negative dilatancy angle would
represent contraction of the specimen. From Table 3, higher dilatancy
4. Results angle of 6.4° was recorded for sandstone while a relatively lower di­
latancy angle of 2.0° was produced for shale spoils at a normal stress of
4.1. Direct shear test results 300 kPa. At 500 kPa, sandstone produced a dilatancy angle of 4.1°
while shale produced a negative dilatancy angle of −0.1°, which was
Direct shear tests were conducted on tunneling spoils of sandstone due to compression of the test specimen. From Table 3, it can be ob­
and shale at normal stresses of 100 kPa, 300 kPa and 500 kPa. Table 3 served that with increasing normal stresses, a reduction in dilatancy
shows the results of peak and residual shear strength along with the angle is noticeable.

Table 3
Comparison of results of peak and residual shear strength for sandstone and shale spoils.
Material Tangent at normal stress (kPa) Dilatancy angle, (ψ) (degree)a Shear stress (τ) (kPa) Apparent cohesion (c′) (kPa) Internal friction angle (ϕ′) (degree)

Peak Residual Peak Residual Peak Residual

Sandstone 100 8.2 97.9 64.8 14.7 4.1 39.4 30.9


300 6.4 241.6 177.7 37.3 11.5 34.8 29.2
500 4.1 383.6 292.1 57.5 18.6 32.8 28.4
Shaleb 100 2.4 90.5 73.8 0.0 0.0 37.3 35.0
300 2.0 236.8 201.2
500 −0.1 395.7 359.5

Note:
a
Measured at occurrence of peak shear stresses.
b
Line of best fit was sufficient to develop M−C failure criterion.

5
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

Sandstone demonstrated nonlinear stress–strain behavior while the well-graded sandstone particles.
shale spoils exhibited a linear stress–strain behavior as shown in Fig. 3
(a) and (b) respectively. If a best-fit MC strength criterion is adopted
4.2. GeoPIV results
conveniently, Choo and Ong (2015) found that an over-estimation of
shear stresses may happen at lower and higher normal stresses while an
Vector plots were produced with respect to the corresponding stages
under-estimation of shear stresses at intermediate normal stresses when
in the four-state shear model (Li and Aydin, 2010). Particle movements
testing non-linear strength materials. Therefore, to overcome such in­
were captured at various stages of shearing for each of the applied
accuracies, De Mello’s (1977) Power Law function to characterize the
normal stresses. The directions of particle movements were represented
non-linear strength behavior of such materials is proposed to be used.
with vectors, while the magnitude of particle displacements were re­
Eq. (5) defines the simplified Power Law function.
presented by the lengths of the vectors. The upper and lower halves of
= A. ( ') B (5) the shear box were easily distinguishible, where the top half consisted
of relatively less movements as compared to the lower half of the shear
where, A is defined as a dimensionless constant which controls the box. This was because only the lower half was displaced during the
magnitude of the power function while B is a dimensionless constant shearing of the test specimens; the top half of the shear box and the
which controls the curvature of the power function (De Mello, 1977). camera were stationery during the tests. During the shearing process,
The power law parameters, A and B for the sandstone specimens are vector plots showed particles moving downwards at the right end of the
1.95 and 0.85 respectively, as shown in Fig. 3 (a). Similar approach shear box while an upward movement was recorded on the left end of
using power law parameters, A and B have been documented by Charles the shear box, as shown by the dotted arrow in Fig. 4. Several re­
and Watts (1980); De Mello (1977); Choo and Ong (2015) and Ong searchers found similar observations during discrete element modelling
et al. (2018). for granular materials (O'Sullivan et al., 2006; Kang et al., 2012;
To assess the arching effect, the MC strength parameters were still Indraratna et al., 2014; Salazar et al., 2015; Peerun et al., 2019). Other
required. Hence, Yang and Yin (2004) proposed a generalized tangen­ activities such as localized dilation and compression were also found in
tial technique that can be used to obtain the tangential friction angle the vector plots.
and tangential cohesion from the nonlinear behavior of the tested Using both the direct shear test and GeoPIV results together, the
specimens in this current study. The generalized tangential method behaviors of the particles along the shear band can be interpreted to­
consists of drawing a tangent to the power law strength envelope at the gether with the specimen strength parameters at different shearing
appropriate normal stress, which in this study corresponded with the in- stages. Localized activities can be observed and linked to its respective
situ overburden pressures at the respective pipe-jacking tunnel inverts. shear stress and volumetric plots. Localized compression could be due
The strength envelopes for sandstone specimens showed non-linearity to possible particle breakages while localized dilation could be due to
and hence the power law function was applied with tangents drawn at interlocking. It is to be mentioned that particle breakage do not form
100 kPa, 300 kPa and 500 kPa normal stresses for peak and residual part of this study.
stresses. This technique was used to interpret friction angle and cohe­ From Table 3, sandstone spoils produced larger apparent cohesion
sion of reconstituted tunneling rock spoils by Choo and Ong (2015), with an increase of normal stress for both peak and residual states. At
Ong and Choo (2016) and Peerun et al. (2019). A best-fit MC line was 100 kPa, sandstone produced an apparent cohesion of 14.7 kPa at peak
used for shale specimen as it showed linear strength behavior. Well- which was increased up to 18.6 kPa for a normal stress of 500 kPa. At
graded sandstone produced relatively higher apparent cohesion due to 500 kPa normal stress, an apparent cohesion of 3.9 times at peak state
its coarse particles as compared to shale, which was comprised of flaky was recorded as opposed to a normal stress of 100 kPa. Figs. 4, 5 and 6
micaceous particles. Sandstone particles consisting of strong quartz show the vector plots for sandstone spoils during Stage 2 shearing at
constituents could have greater interlocking and hence resulted in normal stresses of 100 kPa, 300 kPa and 500 kPa respectively. As per
higher apparent cohesion. However, limited apparent cohesion was this stage, the particles interlock against one another which resulted in
observed for shale spoils. It could be due to minimal interlocking among dilation of the specimen as shown by the vertically inclined vectors in
its particles, which consist of fine-grained clay minerals as compared to Fig. 4. The maximum state of dilation would hence produce a peak

Fig. 3. (a) Non-linear strength envelope of sandstone spoils and (b) M−C linear strength envelope of shale spoils.

6
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

Fig. 4. Dilation due to interlocking activities for sandstone spoils during Stage 2 shearing at normal stresses of 100 kPa.

Fig. 5. Dilation due to interlocking activities for sandstone spoils during Stage 2 shearing at normal stresses of 300 kPa.

Fig. 6. Dilation due to interlocking activities for sandstone spoils during Stage 2 shearing at normal stresses of 500 kPa.

shear stress. With increasing normal stresses of 300 kPa and 500 kPa angle of 4.1° was observed at 500 kPa. Therefore, at higher normal
(Figs. 5 and 6), a reduction in dilation activities is observed respec­ stresses, larger contact areas among the particles would produce more
tively. This is due to the relatively higher normal stresses acting onto interlocking but due to the larger normal load, it would be more dif­
the sandstone particles, producing more contact areas among the par­ ficult for the particles to roll over the surrounding particles. The latter
ticles and making it more difficult for the particles to roll onto one would result in greater resistance, which is demonstrated in terms of
another during the interlocking stage. The volumetric plots in Fig. 7 apparent cohesion. Table 3 also shows with an increase of normal
confirm that the most dilation occurs at 100 kPa (vertical deforma­ stresses, relatively larger apparent cohesion was obtained with a cor­
tion = 0.3 mm) and while the least dilation occurs at 500 kPa normal responding reduction in friction angle. At peak state from 100 kPa to
stress (vertical deformation = 0.18 mm). Furthermore, Table 3 also 500 kPa, an apparent cohesion of 14.7 kPa was increased to 57.5 kPa
validates this important observation where with the highest dilatancy with increasing normal stress while the internal friction angle of 39.4°
angle was recorded at 8.2° for 100 kPa and while the lowest dilatancy was reduced to 32.8°.

7
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

While particle interlocking generate dilation, particle breakages


would result in compression of the specimen. Shale comprises of rough
angular particles which were more prone to particle breakages as it had
relatively weaker mineralogical content consisting of fine-grained clay
minerals. An attempt in studying particle breakages was conducted by
plotting the particle size distributions (PSD) before and after shearing.
For such relatively small specimen volume of 3.34 × 10-4 m3, the small
amount of particle breakages could not be represented on a PSD curve.
Kim and Ha (2014) showed similar results where a large shear box was
used to study the effect of particle size on shear behavior of sand and
gravel using geogrids as reinforcement. The PSD curve before and after
the test showed insignificant change and particle crushing could not be
conclusive. Hence, with the aid of novel vector plots from the GeoPIV
analysis, localized activities could be identified such as dilation and
compression representing particle interlocking and possible breakages,
respectively (Peerun et al. 2019). Figs. 8 and 9 show the vector plots at
Fig. 7. Vertical deformation plot for sandstone spoils at 100 kPa, 300 kPa and
Stage 3 under normal stress of 300 kPa for sandstone and shale
500 kPa normal stresses with solid lines representing shearing Stage 2.

Fig. 8. Vector plots showing localized interlocking for sandstone at Stage 3 and confining pressure of 300 kPa.

8
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

Fig. 9. Vector plots showing localized interlocking for shale at Stage 3 and confining pressure of 300 kPa.

respectively. In contrast to dilation observed in Fig. 8, the shale spe­ shale spoils produced no cohesion as per Table 3 and showed linear
cimen consisted of localized compression which could be due to pos­ strength behavior based on Fig. 3 (b).
sible particle breakages. The non-existent particle interlocking in shale The schematic diagram of Fig. 10 shows the influence of particle
can be concluded from the localized compression and minimal dilation shapes and mineralogy on the shear stress. The well-graded and rela­
in Fig. 9. With minimal interlocking activities, the relatively weaker tively strong sandstone particles consisting low void ratio produced

9
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

Fig. 10. Influence of (a) particle shapes and mineralogy on (b) shear stress and (c) shear strength.

Fig. 11. Representation of change in shear stresses during direct shear test in assessment of arching using Pellet-Beaucour and Kastner (2002) model.

greater inter-particle contact area against one another to produce a tunnel drives traversing sandstone and shale geology. This was
distinct peak shear stress. On the other hand, the weaker angular par­ achieved by using the soil properties obtained from direct shear tests
ticles of shale, which consisted relatively higher void ratio, was com­ and Eqs. (1), (2) and (3) to assess the vertical stresses acting on the pipe
pressed and resulted in a steady shear stress distribution at larger dis­ crown of each drives. The arching effect is defined as the transfer of
placement. Therefore, it is thought that the interlocking particles pressure from a yielding mass of soil onto adjoining stationary parts
produce an apparent cohesion for the non-linear strength behavior of (Terzaghi, 1943). Arching is maintained purely by shearing stresses,
sandstone while the flaky micaceous shale spoils breaks and then slide which is strongly related to the direct shear test. The change in stresses
over one another without further interaction, thus resulting in a linear during DST where the soil in the upper half of the shear box is displaced
MC strength behavior, which subsequently produce a friction angle and along the stationary soil in the lower half of the shear box is a good
zero cohesion. representation of the change in stresses along the vertical boundary
plane of a developing wedge failure as shown schematically in Fig. 11.
Therefore, the shear strength properties obtained from DST can be used
5. Case studies
to quantify the influence of apparent cohesion and friction angle on soil
arching. From Pellet-Beaucour and Kastner (2002) model, φ’ and c’ are
This section discusses the assessment of jacking forces for four

10
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

Fig. 12. Soil profiles for Drive A and B.

influencing factors in defining the vertical soil stress acting on the pipe pipeline, no lubrication was used while an average of 250 L/m were
crown (Eq. (2)) and φ’ in defining the silo width, b (Eq. (3)). The soil used for the second half. The jacking speed was recorded ranging be­
profiles and properties of the four tunnel drives are shown in Figs. 12 tween 10 and 40 mm/min with some occasional spikes. However, this
and 13. produced a steady drive with a measured jacking force of 10.1 kN/m.
Fig. 14 (a) shows the jacking forces, jacking speed and lubrication used
5.1. Drive A for Drive A. Direct shear tests were conducted on sandstone spoils to
obtain the non-linear strength parameters and used to interpret Drive A
Drive A consisted of jacking a concrete pipeline of 65 m and 1.43 passing through the excavated sandstone. The apparent cohesion, cpeak’
diameter at a depth of 9.50 m. Based on soil investigations, the pipeline was 11.9 kPa and friction angle, ϕpeak’ was 40.5°. By inputting these
traversed moderately weathered sandstone. During the first half of the strength parameters in Eq. (2), the relatively strong, rounded particles

11
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

Fig. 13. Soil profiles for Drive C and D.

of sandstone produced a vertical stress, σEV of 15.3 kPa acting on the friction angle and cohesion were obtained based on the effective
pipeline. Vertical soil stress at the pipe crown, σEV lower than zero overburden pressure for each drive.
(negative values) is an indication of the absence of soil pressure acting For Drive A in sandstone, a relatively smaller amount of σEV of
on the jacked sewer pipeline (presence of strong arching) while a po­ 15.3 kPa was found acting on the pipe crown. This may not demon­
sitive value of the vertical soil stress at the pipe crown, σEV would in­ strate strong presence of arching but it shows lower soil pressure acting
dicate soil contact on the pipe (presence of less arching) (Choo and Ong, on the pipe crown as compared to its overburden pressure of 78 kPa.
2015, 2017; Ong et al., 2018). σEV computed for the tunneling drives The significant amount of dilation as shown in Fig. 8 was due to particle
can be found in Table 4 together with the relevant parameters. The soil interlocking amongst the strong and rounded quartz particles,

12
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

Fig. 14. Variations in measured jacking forces, jacking speed, lubricant and back-analysed jacking forces for (a) Drive A, (b) Drive B.

13
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

Table 4
Computation of vertical soil stresses, σEV.
Parameters Symbol Units Drive A Drive B Drive C Drive D

Geology – – Sandstone Shale

Outer pipe diameter De m 1.43 1.43 1.43 1.78


Effective overburden pressure σ'V kPa 78.0 121.7 110 156
Average soil unit weight γ kN/m3 18.21 19.19 18.64 18.44
Soil cover from ground to pipe crown h M 9.50 13.5 12.5 18.0
Soil friction angle (based on Power Law) ϕ(a) degree 40.5 28.6 37.3 37.3
Soil cohesion (based on Power Law) C(b) kPa 11.9 17.3 0.0 0.0
Lateral earth pressure coefficient K – 1.0 1 1 1
Influencing soil width above pipe b m 2.70 2.8 2.8 3.5
Thrust coefficient K2 – 0.3 0.3 0.3 0.3
Jacking force measured (F/L) JF kN/m 10.1 18.9 29.0 28.5
Calculated vertical stress acting on pipe crown from Eq. (2) σEV kN/m2 15.3 12.1 34.8 42.9
Calculated frictional coefficient from Eq. (1) μavg – 0.122 0.251 0.207 0.132

producing relatively higher shear stress and apparent cohesion as listed computed to be 34.8 kPa which is 2.3 times of Drive A and 2.9 times of
in Table 3. From Table 4, the higher frictional angle of 40.5° and co­ Drive B. Hence, a relatively greater soil pressure was acting on the pipe
hesion of 11.9 kPa produced a σEV of 15.3 kPa which made it possible to which then required a 29.0 kN/m jacking force to tunnel through shale.
jack through sandstone steadily at an average jacking force of 10.1 kN/ Again, a relatively larger jacking force of 2.9 times of Drive A and 1.5
m, an average jacking speed of 29.2 mm/min and requiring sig­ times of Drive B, was required but with a relatively much slower
nificantly lower lubrication (250 L/m). average jacking speed of 10.4 mm/min. An attempt to reduce the
jacking force and to speed up the process was made by increasing the
5.2. Drive B amount of lubrication as shown in Fig. 15 (a). In total, up to 2500 L/m
were eventually used for this drive. However, the effectiveness of the
The second drive consisted of a 160 m long and 1.43 m diameter lubricant is doubtful due to the continous high jacking force and re­
concrete pipeline which traversed a moderately strong light grey duced jacking speed. This could be due to loss of lubrication to the
slightly weathered sandstone at a depth of 13.5 m. The soil profile along surrounding fissures/cracks of shale as discussed in detail by Choo and
with its properties are found in Fig. 12. From the 13.5 m of soil cover to Ong (2015).
the pipe crown, the tangential power law MC parameters of 28.6°
frictional angle and 17.3 kPa cohesion were obtained for a soil pressure
of 121.7 kPa. By incorporating these parameters into Eq. (2), a σEV of 5.4. Drive D
12.1 kPa was found acting on the pipe crown. Similar to Drive A, a
lower soil pressure of 12.1 kPa was found acting on the pipe crown as This drive consisted of a relatively short tunnel span of 30 m only
compared to its overburden pressure of 121.7 kPa. The significantly low traversing very weak grey highly weathered shale at a depth of 18 m.
soil pressure acting on the pipe crown again is a result of relatively The weak carboneous angular shale spoils produced a frictional angle of
significantly high frictional angle and cohesion values which are the 37.3° with no cohesion due to localized compression as discussed in
product of interparticle activities as discussed for Drive A. As such, a Drive C, producing σEV of 42.9 kPa acting on the pipe crown. Again, this
jacking force of 18.9 kN/m was required to complete the 160 m drive. significantly larger soil pressure (2.8 times of Drive A and 3.5 times of
This was achieved at an average jacking speed of up to 22.4 mm/min Drive B) required a larger jacking force of 28.5 kN/m to jack through
and required minimal constant lubricants of 250 L/m as shown in the challenging shale geology. This is 2.8 and 1.5 times greater than for
Fig. 14 (b). Drives A and B in sandstone, respectively. An average jacking speed of
18.9 mm/min and lubrication of up to 500 L/m were recorded for the
5.3. Drive C short drive in shale (Fig. 15 (b)). The overall comparison between the
sandstone and shale drives showed distinct differences in magnitudes in
Drive C traverses a very weak grey highly weathered shale for a terms of jacking force, jacking speed and lubricant used. Drives A and B
span of 140 m at a depth of 12.5 m. The poorly graded shale spoils were required lower jacking forces to tunnel through sandstone at a higher
subjected to direct shear tests and produced a frictional angle of 37.3° jacking speed and significantly lower amount of lubricants used as
with no cohesion. When compared to sandstone, shale specimens pro­ compared to Drives C and D in shale. Such assessment was achieved by
duced localized compression which is due to suspected particle analyzing σEV acting on the pipe crown, which was found to be de­
breakages as shown in Fig. 9, resulting in lower shear stresses and no pendent on the soil frictional angle and cohesion in this paper.
cohesion. From the MC parameters, σEV acting on the pipe crown was

14
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

Fig. 15. Variations in measured jacking forces, jacking speed, lubricant and back-analysed jacking forces for (a) Drive C and (b) Drive D.

6. Conclusions calibrated GeoPIV technique. The interpretation of shear strength re­


sults along with the GeoPIV results have provided a better insight on
This study focuses on the assessment of the arching phenomenon the behaviors of the particles during shearing. Furthermore, the out­
with the aid of particle behavior such as localized compression or di­ comes were used to assess the arching phenomenon during pipe-jacking
lation in a specially fabricated and enhanced transparent shear box, as works based on established case studies. The following observations
evidenced by the series of images reliably analyzed using the well- were made:

15
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

1) materials with stronger particles mineralogy are prone to greater Choo, C.S., Ong, D.E.L., 2015. Evaluation of Pipe-Jacking Forces Based on Direct Shear
dilation as compared to weaker matrices particles content; Testing of Reconstituted Tunneling Rock Spoils. J. Geotech. Geoenviron. Eng. 141
(10). https://doi.org/10.1061/(ASCE)GT.1943-5606.0001348.
2) Angular particles would produce a sample with higher void ratio Choo, C.S., Ong, D.E.L., 2017. Impact of highly weathered geology on pipe-jacking forces.
and hence are subjected to greater compression as compared to rounded Geotech. Res. 4 (2), 94–106. https://doi.org/10.1680/jgere.16.00022.
shape particles; Choo, C.S., Ong, D.E.L., 2020. Assessment of non-linear rock strength parameters for the
estimation of pipe-jacking forces. Part 2. Numerical modeling. Eng. Geol. 265,
3) Particle interlocking will produce greater dilation and resulting 105405. https://doi.org/10.1016/j.enggeo.2019.105405.
in higher apparent cohesion; De Mello, V.F.B., 1977. Reflections on design decisions of practical significance to em­
4) Specimen compression due to possible particles breakages will bankment dams. Géotechnique 27 (3), 281–355. https://doi.org/10.1680/geot.1977.
27.3.281.
reduce the peak shear stress; DeGregorio, V.B., 1990. Loading Systems, Sample Preparation, and Liquefaction. J.
5) At higher normal stresses, increase in interlocking activities will Geotech. Eng. 116 (5), 805–821. https://doi.org/10.1061/(ASCE)0733-9410(1990)
produce higher apparent cohesion with a corresponding reduction in 116:5(805).
DeJong, J.T., Westgate, Z.J., 2009. Role of Initial State, Material Properties, and
friction angle;
Confinement Condition on Local and Global Soil-Structure Interface Behavior. J.
6) Greater apparent cohesion will result in a reduction of over­ Geotech. Geoenviron. Eng. 135 (11), 1646–1660. https://doi.org/10.1061/(ASCE)
burden pressure or vertical soil stresses acting on the jacked pipe, hence 1090-0241(2009)135:11(1646).
demonstrating effect of arching. Therefore, as illustrated by the fair Fukuoka, H., Sassa, K., Wang, G., Sasaki, R., 2006. Observation of shear zone develop­
ment in ring-shear apparatus with a transparent shear box. Landslides 3 (3), 239–251.
agreement between the laboratory results and the well-established field https://doi.org/10.1007/s10346-006-0043-2.
case studies, by being able to express the significance of arching for Grognet, M., 2011. The boundary conditions in direct simple shear tests, development for
different geological settings in the field, engineers can estimate the peat testing at low vertical stress. MSc thesis thesis. Delft University of Technology.
Indraratna, B., Ngo, N.T., Rujikiatkamjorn, C., Vinod, J.S., 2014. Behavior of Fresh and
impact of σEV acting on the jacked pipe. This knowledge would enable Fouled Railway Ballast Subjected to Direct Shear Testing: Discrete Element
engineers to not only plan the jackable length of the pipe-jacking drive Simulation. Int. J. Geomech. 14 (1), 34–44. https://doi.org/10.1061/(ASCE)GM.
more efficiently, but also being able to estimate quantitatively the 1943-5622.0000264.
Jackson, P., 2010. An Investigation into the Deformation Behaviour of Geosynthetic
jacking force. Therefore, by incorporating the knowledge of the arching Reinforced Soil Walls under Seismic Loading. MEng Thesis. University of
phenomenon into soil-structure interaction, in this case, pipe-jacking Canterbury.
works in fractured rock mass, a more efficient and economical design Jing, X.-Y., Zhou, W.-H., Zhu, H.-X., Yin, Z.-Y., Li, Y., 2017. Analysis of soil-structural
interface behavior using three-dimensional DEM simulations. Int. J. Numer. Anal.
can be subsequently achieved. Methods Geomech. 42 (2), 339–357. https://doi.org/10.1002/nag.2745.
Kang, D.H., Lee, J.H., Choo, J., Yun, T.S., 2012. Pore directivity of soils subjected to
CRediT authorship contribution statement shearing: numerical simulation and image processing. GeoCongress 2012,
2342–2351. https://doi.org/10.1061/9780784412121.240.
Kelly, P., 2014. Soil Structure Interaction and Group Mechanics of Vibrated Stone Column
M.I. Peerun: Data curation, Formal analysis, Validation, Software, Foundations. PhD Thesis. University of Sheffield.
Writing - original draft. D.E.L. Ong: Conceptualization, Methodology, Khondoker, M. T. H., Yi, Y., and Bayat, A., 2016. Comparison of different methods for
Writing - review & editing, Supervision, Project administration, normal stres calculation during pipe jacking/microtunnelling. Int. J. of Geo. Mech.
https://doi.org/10.1080/19386362.2016.1147193.
Funding acquisition. C.S. Choo: Data curation, Methodology, Kim, D., and Ha, S., 2014. Effects of particle size on the shear behavior of coarse grained
Resources, Writing - review & editing. W.C. Cheng: Writing - review & soils reinforced with geogrid. Materials, 7963–979. https://doi.org/10.3390/
editing. ma7020963.
Li, Y.R., Aydin, A., 2010. Behavior of rounded granular materials in direct shear:
Mechanisms and quantification of fluctuations. Eng. Geol. 115 (1-2), 96–104. https://
Declaration of Competing Interest doi.org/10.1016/j.enggeo.2010.06.008.
Mehdizadeh, A., Disfani, M.M., Evans, R., Arulrajah, A., Ong, D.E.L., 2016. Discussion of
“Development of an internal camera-based volume determination system for triaxial
The authors declare that they have no known competing financial testing”. Geotech. Test. J. 39 (1), 165–168. https://doi.org/10.1520/GTJ20150153.
interests or personal relationships that could have appeared to influ­ Mehdizadeh, A., Disfani, M.M., Evans, R., Arulrajah, A., Ong, D.E.L., 2017. Mechanical
Consequences of Suffusion on Undrained Behaviour of a Gap-Graded Cohesionless
ence the work reported in this paper. Soil - An Experimental Approach. Geotech. Test. J. 40 (6), 1026–1042. https://doi.
org/10.1520/GTJ20160145.
Acknowledgement Misra, A., Roberts, L.A., Najafi, M., 2008. Probabilistic soil-structure interaction model for
pipe-jacking force analysis, Proceedings of Pipelines 2008: Pipeline Asset
Management: Maximizing Performance of our Pipeline Infrastructure. American
The authors would like to thank David White for providing the Society of Civil Engineers, New York, pp. 1–10.
GeoPIV software and their grateful appreciation to Hock Seng Lee Bhd O'Sullivan, C., Bray, J.D., Cui, L., 2006. Experimental validation of particle-based discrete
and Jurutera Jasa (Sarawak) Sdn Bhd in providing field data for this element methods. GeoCongress 1–18. https://doi.org/10.1061/40803(187)5.
Oda, M., Kunishi, J., 1974. Microscopic deformation mechanism of granular material in
study. simple shear. Japan Soc. Soil Mech. Found. 14 (4), 25–38.
Olson, M.P., 2013. Pilot tube microtunneling: instrumentation and monitoring for jacking
References force and productivity analysis, MSc thesis, Arizona State University, Tempe, AZ.
Olson, M.P., Ariaratnam, S.T., Lueke, J.S., 2015. Jacking force and productivity analysis
of pilot tube microtunneling installations, J. Pipeline Syst. Eng. Pract., 04015018-1.
ASTM, 2000. Standard practice for classification of soils for engineering purposes (unified Ong, D.E.L., Leung, C.F., Chow, Y.K., 2003. Piles subject to excavation-induced soil
soil classification system). D2487-00, West Conshohocken, PA. movement in clay. Proc. 13th European Conference on Soil Mechanics and
ASTM, 2011. Standard Test Method for Direct Shear Test of Soils Under Consolidated Geotechnical Engineering, Prague, Czech Republic, Vol. 2, pp. 777–782.
Drained Conditions, D3080M-11, West Conshohocken, PA. Ong, D.E.L., Yang, D.Q., Phang, S.K., 2006. Comparison of finite element modelling of a
Bolton, M.D., 1986. The strength and dilatancy of sands. Géotechnique 36 (1), 65–78. deep excavation using SAGE-CRISP and PLAXIS. Int. Conf. on Deep Excavations, 28-
Charles, J.A., Watts, K.S., 1980. The influence of confining pressure on the shear strength 30 June 2006, Singapore, pp. 51–64.
of compacted rockfill. Géotechnique 30 (4), 353–367. https://doi.org/10.1680/geot. Ong, D.E.L., Choo, C.S., 2011. Sustainable Bored Pile Construction in Erratic Phyllite.
1980.30.4.353. ASEAN-Australian Engineering Congress. Kuching, Malaysia (ISBN 978-967-10485),
Chapman, D.N., Ichioka, Y., 1999. Prediction of jacking forces for microtunnelling op­ July 2011, pp. 30–45.
erations. Tunn. Undergr. Space Technol. 14, 31–41. Ong, D.E.L., Choo, C.S., 2016. Back-analysis and finite element modeling of jacking forces
Cheng, W., Ni, J.C., Arulrajah, A., Huang, H., 2018. A simple approach for characterising in weathered rocks. Tunn. Undergr. Space Technol. 51, 1–10. https://doi.org/10.
tunnel bore conditions based upon pipe-jacking data. Tunn. Undergr. Space Technol. 1016/j.tust.2015.10.014.
71 (2018), 494–504. https://doi.org/10.1016/j.tust.2017.10.002. Ong, D.E.L., Choo, C.S., 2018. Assessment of non-linear rock strength parameters for the
Cheng, W.-C., Ni, J.C., Huang, H.-W., Shen, J.S., 2018b. The use of tunnelling parameters estimation of pipe-jacking forces. Part 1. Direct shear testing and backanlaysis. Eng.
and spoil characteristics to assess soil types: a case study from alluvial deposits at a Geol. 244 (2018), 159–172. https://doi.org/10.1016/j.enggeo.2018.07.013.
pipejacking project site. Bull. Eng. Geol. Environ. 78 (2019), 2933–2942. https://doi. Ong, D.E.L., Sim, Y.S., Leung, C.F., 2018. Performance of Field and Numerical Back-
org/10.1007/s10064-018-1288-4. Analysis of Floating Stone Columns in Soft Clay Considering the Influence of
Cheng, W.-C., Wang, L., Xue, Z.-F., Ni, J.C., Rahman, M.M., Arulrajah, A., 2019. Dilatancy. Int. J. Geomech. 18 (10). https://doi.org/10.1061/(ASCE)GM.1943-5622.
Lubrication performance of pipejacking in soft alluvial deposits. Tunn. Undergr. 0001261.
Space Technol. 91, 102991. https://doi.org/10.1016/j.tust.2019.102991. Osumi, T., 2000. Calculating jacking forces for pipe jacking methods. No-Dig Int. Res.

16
M.I. Peerun, et al. Tunnelling and Underground Space Technology 106 (2020) 103610

40–42. 001.
Peerun, M.I, Ong, D.E.L., Choo, C.S., 2019. Interpretation of Geomaterial Behavior during Terzaghi, K., 1936. Stress distribution in dry and in saturated sand above a yielding trap
Shearing Aided by PIV Technology. J. Mater. Civil Eng., 31(9). https://doi.org/10. door. Proc., 1st Int. Conf. on Soil Mechanics and Foundation Engineering, Harvard
1061/(ASCE) MT.1943-5533.0002834. Univ., Cambridge, MA, 307–311.
Pellet-Beaucour, A.-L., Kastner, R., 2002. Experimental and analytical study of friction Terzaghi, K., 1943. Conditions for Shear Failure in Ideal Soils. Theoretical Soil Mechanics,
forces during microtunneling operations. Tunn. Undergr. Space Technol. 17 (1), John Wiley & Sons, Inc.
83–97. Wafid, M.A., Sassa, K., Fukuoka, H., Wang, G.H., 2004. Evolution of shear-zone structure
Pfeffer, W.T., 2014, Sedimentary rocks, CVN 3698 Engineering geology spring, INSTAAR in undrained ring-shear tests. Landslide 1 (2), 101–112.
Dept. Civil. Env. Architectural Eng,University of Colorado. Wang, G.H., Sassa, K., 2002. Post mobility of saturated sands in undrained load-controlled
Rahjoo, S., Najafi, M., Williammee, R., Khankarli, G., 2012. Comparison of jacking load ring shear tests. Can. Geotech. J. 3\9, 4, 821–837.
models for trenchless pipe jacking, Pipelines 2012: Innovations in design, construc­ Wang, H., Zhou, W., Yin, Z., Jie, X., 2019. Effect of grain size distribution of sandy soil on
tion, operations, and maintenance, doing more with less. American Society of Civil shearing behaviors at soil-structure interface. J. Mater. Viv. Eng. (ASCE) 31 (10).
Engineers, New York, pp. 1507–1520. https://doi.org/10.1061/(ASCE)MT.1943-5533.0002880.
Salazar, A., Sáez, E., Pardo, G., 2015. Modeling the direct shear test of a coarse sand using White, D.J., Take, W.A., 2002. GeoPIV: Particle Image Velocimetry (PIV) software for use
the 3D Discrete Element Method with a rolling friction model. Comput. Geotech. 67, in geotechnical testing, Eng. Dept. Cambridge University.
83–93. https://doi.org/10.1016/j.compgeo.2015.02.017. Yang, X.L., Yin, J.H., 2004. Slope stability analysis with nonlinear failure criterion. J. Eng.
Shen, S.-L., Cui, Q.-L., Ho, C.-E., Xu, Y.-S., 2016. Ground Response to Multiple Parallel Mech., 130, 3, 267–273. https://doi.org/10.1061/(ASCE)0733-9399(2004)
Microtunneling Operations in Cemented Silty Clay and Sand. J. Geotech. Geoenviron. 130:3(267).
Eng. 142 (5). https://doi.org/10.1061/(ASCE)GT.1943-5606.0001441. Zhang, H., Zhang, P., Zhou, W., Dong, S., Ma, B., 2016. A new model to predict soil
Shimizu, M., 1997. Strain fields in direct shear box tests on a metal-rods model of pressure acting on deep burial jacked pipes. Tunn. Undergr. Space Technol. 60,
granular soils. In: Asaoka, Adachi, Oka. Deform. Progr. Failure Geomech., 151–156. 183–196. https://doi.org/10.1016/j.tust.2016.09.005.
Shou, K., Yen, J., Liu, M., 2010. On the frictional property of lubricants and its impact on Zhou, W.-H., Jing, X.-Y., Yin, Z.-Y., Geng, X., 2019. Effects of particle sphericity and
jacking force and soil–pipe interaction of pipe-jacking. Tunn. Undergr. Space initial fabric on the shearing behavior of soil–rough structural interface. Acta
Technol. 25 (4), 469–477. https://doi.org/10.1016/j.tust.2010.02.009. Geotech. 14 (6), 1699–1716. https://doi.org/10.1007/s11440-019-00781-2.
Staheli, K., 2006. Jacking force prediction: an interface friction approach based on pipe Zhu, H.-X., Zhou, W.-H., Jing, X.-Y., Yin, Z.-Y., 2019. Observations on fabric evolution to
surface roughness. PhD thesis. Georgia Institute of Technology. a common micromechanical state at the soil‐structure interface. Int. J. Numer. Anal.
Stein, D., 2005. Trenchless technology for installation of cables and pipelines. Prof. Dr.- Methods Geomech. 43 (15), 2449–2470. https://doi.org/10.1002/nag.2989.
Ing. Stein & Partner GmbH, Bochum, Germany. Zhu, X., Gong, W., 2014. Study on soil arching in the cushion of composite foundation.
Sterling, R.L., 2020. Developments and research directions in pipe jacking and micro­ Soil Behaviour and Geomechanics, Geo-Shanghai 2014, ASCE, https://doi.org/10.
tunneling. Undergr. Space 5 (1), 1–19. https://doi.org/10.1016/j.undsp.2018.09. 1061/9780784413388.058.

17

You might also like