You are on page 1of 11

Tunnelling and Underground Space Technology 71 (2018) 494–504

Contents lists available at ScienceDirect

Tunnelling and Underground Space Technology


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

A simple approach for characterising tunnel bore conditions based upon MARK
pipe-jacking data

Wen-Chieh Chenga, , James C. Nib, Arul Arulrajahc, Hui-Wen Huangb
a
Institute of Tunnel and Underground Structure Engineering, School of Civil Engineering, Xi’an University of Architecture and Technology, Shaanxi 710055, China
b
Department of Civil Engineering, National Taipei University of Technology, Taipei 10608, Taiwan, ROC
c
Department of Civil and Construction Engineering, Swinburne University of Technology, Victoria 3122, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: There are several well-established jacking force models available for determining the jacking loads. However,
Normal contact pressure their ability to characterise the tunnel bore conditions is limited. A simple approach to characterise the tunnel
Back-analysed frictional coefficient bore conditions is proposed and applied to a case study where four sewer pipelines of the Shulin district sewer
Soil arching network in Taipei County, Taiwan were constructed to verify its validity. In this paper, four jacking force models
are reviewed. Based upon the given soil properties and pipe dimensions as well as the pipe buried depth, the
normal contact pressure (σ′) in each jacking force model and the measured frictional stress (τ) in each baseline
section are utilised for back-analysis of the frictional coefficient (μavg). The μavg values outside the range of
0.1–0.3 recommended for lubricated drives can be attributed to the increasing pipe friction resulting from ex-
cessive pipe deviation or ground closure or due to the gravel formation not being long enough to establish lower
face resistance or total jacking load. JMTA (Japan Microtunnelling Association) has indicated a further potential
use in assessment of the interface performance during pipe-jacking works.

1. Introduction Shulin district sewer network in Taipei County, Taiwan was analysed
using the baseline technique. The specific objectives of this study are (i)
Jacking force prediction highly correlates with the design and se- to review four well-established jacking force models, (ii) to propose a
lection of a pipe-jacking system (Chapman and Ichioka, 1999; Sofianos simple approach that is capable of characterising the tunnel bore con-
et al., 2004; Khazaei et al., 2004; Staheli, 2006; Barla et al., 2006; Shou ditions, and (iii) to determine the tunnel bore conditions by comparing
et al., 2010; Rahjoo et al., 2012; Choo and Ong, 2015; Cui et al., 2015; the back-analysed frictional coefficients to the recommended values.
Yen and Shou, 2015; Shen et al., 2016; Wham et al., 2016; Cheng et al.,
2017a,b). There are many well-established jacking force models avail- 2. Jacking force models
able for guiding to the provision of jacking capacity (Rogers and Yonan,
1992; Kastner et al., 1996; Pellet-Beacour and Kastner, 2002; Rahjoo The four well-established jacking force models that were studied
et al., 2012). However, their ability to characterise the tunnel bore are: (i) Japan Microtunnelling Association (JMTA, 2000), hereinafter,
conditions is limited. As tunnelling through the permeable ground, referred to as JMTA, (ii) Ma Baosong (Ma, 2008), (iii) Shimada and
ground closure can be regarded as one of the main factors to lead to Matsui (Shimada and Matsui, 1998), and (iv) Pellet-Beaucour and
some difficulties in estimating the jacking loads. Jet-grouting tech- Kastner (Pellet-Beacour and Kastner, 2002). In the JMTA model, the
nology may be used to enhance the shear strength and watertightness of resistance at the cutting wheel, F0, is theoretically between active and
ground prior to tunnel excavation (Ni and Cheng, 2011a, 2014; Shen passive earth pressure and can be empirically correlated to the blow
et al., 2017). Additionally, misalignment stands for the angular devia- count N-value (JMTA, 2000). The resistance along the pipe string, Fs,
tion between the central axes of successive pipes, and severe mis- considers two components; one from the friction at the pipe-soil in-
alignment can lead to a significant increase in the friction resistance terface and the other from the pipe self-weight induced contact pres-
causing an increment of the jacking loads (Norris, 1992; Ni and Cheng, sure. In the Ma Baosong model, the face resistance, F0, has been as-
2011b, 2012a; Kou et al., 2015; Namli and Guler, 2017). sumed to be equal to the K0 earth pressure against the cutting wheel,
In this study, a pipe-jacking data from four sewer pipelines of the while the friction resistance, Fs, takes the earth pressure acting upon the


Corresponding author.
E-mail addresses: s2428030@gmail.com, w-c.cheng@xauat.edu.cn (W.-C. Cheng).

http://dx.doi.org/10.1016/j.tust.2017.10.002
Received 21 June 2017; Received in revised form 10 October 2017; Accepted 13 October 2017
Available online 02 November 2017
0886-7798/ © 2017 Elsevier Ltd. All rights reserved.
W.-C. Cheng et al.

Table 1
Review of jacking force models.

Jacking force Japan Microtunnelling Association Ma Baosong (Ma, 2008) Shimada and Matsui (Shimada and Matsui, Pellet-Beaucour and Kastner (Pellet-Beacour and Kastner, 2002)
model (JMTA) (JMTA, 2000) 1998)

Face F0 = 10 × 1.32π × De × N F0 = 1/4 × π × De2 × [K 0 Σ(γi ′hi ) + γw hw] where K 0 = F0 = Pw × A where Pw = slurry pressure, A = F0= initial jacking or first load
resis- where De = outer pipe diameter, N coefficient of static soil pressure, hi = thickness of the soil in area of the tunnel face
tance = blow count N value of soil ith layer, γi ′ = density of the soil in ith layer, hw = distance of
(F0)
the groundwater table to the pipe axis, and γw = density of
water
Friction Fs = π × De × τ × L + ω × μ × L Fs = K × [μ × (2PV + 2PH + PB )] where Fs = [p × b × μ1 + Pw × μ 2 × (πD2−b)] × L Fs = μ × L × De × (π /2) × [(σEV + γDe /2) + K2 (σEV + γDe /2)]
resis- where τ = c′ + σ ′tanδ , PV = Kp × γ ′ × h × De × L , where b (contact width between the pipe and where K2 = thrust coefficient of soil arching acting on the pipe,
tance σ′ = γ′ × (De /2)/tanϕ′, c′ = soil bore) = 1.6× (Pu × kd × Ce ) for which with a suggested value of 0.3 (French Society for Trenchless
PH = γ ′ × (h + De /2) × De × L × tan2 (45°−ϕ′/2) , PB = w× L ,
(Fs) cohesion, σ ′ = normal contact Technology, 2006)
K = factor of safety (normally 1.2 is used), PV = vertical Pu = (D12−D22) × γc × 0.25 × π where D1 =

495
pressure acting upon the pipe, δ = pressure of soil above the top of the pipe, PH = lateral earth diameter of bore, D2 = outer diameter of pipe,
angle of wall friction in plane of pressure, PB = total weight of planned jacking pipeline, and γc = concrete density, kd = (D1 × D2) /(D1−D2) ,
sliding, γ ′ = soil density, L = pipe Kp = vertical earth pressure coefficient
Ce = (1−n12) /E1 + (1−n22) /E2 where n1 =
string length, ω = pipe self-weight,
poisson’s ratio of soil, n2 = poisson’s ratio of
and μ = frictional coefficient
concrete pipe, p (contact stress acting on the
pipe bottom) = 2Pu × [1−(x 2 / a2)]0.5 /(π × a) for
which a = b/2 and x = distance to either side of
centerline of the area of contact
Normal σ ′ = γ ′ × (De /2)/tanϕ′ + ω/(πDe ) De 2 ϕ′ ω σ ′ = 2Pu × [1−(x 2 / a2)]0.5 /(π × a) 2×C
σ ′ = Kp × γ ′ × h + γ ′ × h + ) + b × ⎛γ − ⎞
contact
( 2 ) × tan (45°− 2 2 × De ⎝ b ⎠ −2 × K × tanδ ×
× ⎛⎜1−e ⎟
σEV = ( hb ) ⎞ where b (influencing
2 × K × tanδ
pressure ⎝ ⎠
(σ′) π ϕ
silo width) = De × ⎡1 + 2tan −
⎣ ( 4 2 ) ⎤⎦
Tunnelling and Underground Space Technology 71 (2018) 494–504
W.-C. Cheng et al. Tunnelling and Underground Space Technology 71 (2018) 494–504

Table 2
Empirical parameters used in each standard.

Empirical parameters Initial Terzaghi’s definition GB 50332 standard ATV A 161 standard ASTM F 1962 standard BS EN 1594 standard
(Japan) (China) (German) (North America) (British)

b (Silo width): m De × [1 + 2tan(45° − ϕ/2)] De × [1 + tan(45° − ϕ/ De × 3^0.5 1.5De De × [1 + 2tan(45° − ϕ/


2)] 2)]
δ (Angle of wall friction in plane ϕ 30° ϕ/2 ϕ/2 ϕ
of shear): deg.
K (Soil pressure coefficient) 1 tan (45 − ϕ/2)
2
0.5 tan2(45 − ϕ/2) 1 − sinϕ

pipe vertically and horizontally, respectively (Ma, 2008). The face re- lead to some difficulties because they can vary largely from one author
sistance, F0, is expressed as the initial thrust which depends upon the to another. Table 2 lists the empirical parameters adopted in five
slurry pressure. If pipes cannot be maintained to float in the lubricant, standards, (1) the initial Terzaghi’s definition used in Japan (Terzaghi,
the friction resistance, Fs, can be calculated from the elastic solution for 1951), (2) the Chinese national standard GB 50332 (GB 50332-02,
a solid elastic cylinder resting in a cylindrical cavity (Roark and Young, 2002), (3) the German standard ATV A 161 (German ATV rules and
1976; Milligan and Norris, 1996; Shimada and Matsui, 1998). standards, 1990), (4) the American standard ASTM F 1962 (ASTM,
In the Pellet-Beaucour and Kastner model, the initial jacking or first 2011), and (5) the British standard BS EN 1594 (BS, 2009). Their
load is considered to be the face load, F0. The friction resistance, Fs, can comparison shall be discussed as follows.
be calculated based upon the vertical soil stress on the pipe crown (σEV),
which is based on the active trap-door experiment (Terzaghi, 1936) and
represents the difference equation solution of the limit equilibrium of a 3. An approach for characterising tunnel bore conditions
horizontal ground slice subjected to shear stress along the sliding planes
(Terzaghi, 1943). Cheng et al. (2017c) had conducted an investigation An approach proposed in this paper to characterise the tunnel bore
into the effect of the arching phenomenon on the lateral retaining conditions is detailed as follows: (i) establish the baseline of jacking
structure displacement. Table 1 summarises the formulae for calcu- forces, (ii) calculate the normal contact pressure based upon the given
lating the F0 and Fs as well as normal contact pressure (σ′) values, re- soil properties, pipe dimensions, and pipe buried depth, (iii) extract the
spectively, in each model. The σEV value is not only affected by some measured frictional stress in each baseline section, (iv) back-analyse the
physical parameters, i.e., h, C, and γ, but also by empirical parameters, frictional coefficient (μavg) in each baseline section, and (v) determine
i.e., K, b, and δ. These physical parameters can be determined with the tunnel bore conditions by comparing the back-analysed μavg values
some accuracy. However, determining these empirical parameters may to recommended values for lubricated drives.

BH-20 Fig. 1. Location of geological boreholes, geological profile and design


Drive C alignment.
Drive D

BH-03
BH-05
Drive B BH-04
LEGEND
Groundwater level
0 100 m Geological borehole
Residential building
Drive A
Surface backfill
BH-06
Gravelly sand layer
Silty sand layer

14
BH-20
BH-06 EL 11.879m BH-04
12 BH-05
EL 11.120m EL 10.821m
EL 10.866m BH-03
10 EL 9.335m

8
Elevation (m)

0
?
?
-2 ?
? Borehole log
? unavailable ?
-4 ? ?

-6

496
W.-C. Cheng et al. Tunnelling and Underground Space Technology 71 (2018) 494–504

Gravel Sand Silt Fig. 2. Particle size distribution of tunnelling spoils for all pipe-jacking
Coarse Medium Fine Coarse Medium Fine Coarse Medium Fine Clay drives.
100
90 Drive A (GP)
Drive A (GP-GC)
Percent finer by weight (%)

80
Drive A (SW-SM)
70 Drive B (GP)
Drive B (GP)
60
Drive B (SP-SM)
50 Drive B (SP-SC)
Drive C (SP)
40 Drive D (SC)
30 Drive D (GW)

20
10
0
10 1 0.1 0.01 0.001
Particle diameter (mm)

3.1. Project geological conditions total jacking load, with the exception of the second section of gravel at
Drive C and the second section of silty gravel or silty sand at Drive D
Five 15-m deep geological boreholes drilled from a backfill layer where the excessive pipe deviation determined the baseline section
through a silty sand layer into a dense to very dense, poorly-graded to establishment. The measured frictional stress (τ) in each baseline sec-
well graded sand and gravel layer were constructed, as shown in Fig. 1 tion and the normal contact pressure (σ′) in each jacking force model
(Ni et al., 2011, 2013; Ni and Cheng, 2012b). Despite the fine soil not would be utilised for back-analysis of the frictional coefficient (μavg).
being a major component of clayey gravel or clayey sand, the perme- The back-analysed μavg helped not only to assess both the soil-concrete
ability and shear strength of these soil types are controlled by the clay and lubricant-concrete interfaces performance but also to characterise
particles as a sufficient portion of clays can effectively separate each the tunnel bore.
sand particle (D’Appolonia, 1980). The piezometric level is at a depth of In Drive A, the slurry shield spanned 73 m from an initial 5 m sec-
4.8 m, as shown in Fig. 1. tion of gravel through a 62 m section of fine soil governed gravel or
sand deposit and a further 3 m section of gravel into a final 3 m section
3.2. Pipe-jacking works of clay at a depth of 10.3 m (Fig. 3). The jacking force increased from
931 kN at 5 m distance to 1543 kN at 31 m distance while jacking into
The slurry pipe-jacking method was adopted to carry out the ex- the fine soil governed gravel or sand deposit from the gravel formation.
cavations of four pipe-jacking drives in the Shulin district in Taipei Then the jacking force dropped to 1029 kN, the same jacking force
County, Taiwan. The 2.65-m long shield head at 95.1 kN is equipped between 6 and 13 m distance, at 67 m distance while returning to the
with a 1.46-m long articulated steering section functioned by four gravel formation. While jacking into the fine soil governed gravel or
steering cylinders. The 0.92-m long trailing can at 19.6 kN is bolted on sand deposit, the roller discs of the shield might sink into the cutting
to the shield head. During pipe-jacking works, bentonite slurry with a face, leading to full contact between the cutting wheel and the ground,
unit weight of 10.8 kN/m3 was utilised to maintain the stability of the thereby producing additional face resistance due to the increasing
cutting face and tunnel bore and also to pump tunnelling spoils to de- contact area. The local variation of jacking force would give away when
cantation chambers which remove coarse particles from the tunnelling driving back into the coarse soil from the fine soil governed gravel or
spoils through a slurry circulation system. The tunnelling spoils col- sand deposit.
lected from the decantation chambers were used to prepare samples of The increases in the jacking force between 1 and 5 m distance and
the particle size distribution (PSD) analysis based on testing standard between 67 and 70 m distance were most likely due to the pipe frictions
ASTM D422-63 (ASTM, 2002). The PSD curves of the samples are incurred while jacking in the two sections of gravel. Thus, the baseline
shown in Fig. 2. A 1-m long, 1200-mm internal diameter reinforced of jacking forces should consist of three sections, the first 5 m section,
concrete pipe, with a thickness of 120 mm, was adopted in this project. 5–67 m section, and final 67–70 m section. The jacking force in the first
The tunnel bore was excavated using the 1500-mm diameter cutting baseline section averaged 36.8 kN/m, leading to the frictional stress of
wheel. An overcut annulus of 30 mm was established using the smaller 8.1 kPa. The average jacking force in the following two baseline sec-
concrete pipe of 1440 mm in diameter. A special bentonite lubricant tions measured at 1.6 and 32.7 kN/m, respectively, corresponding to
with 2% polymer was utilised to mitigate the impacts of loss of lu- the frictional stresses of 0.4 and 7.2 kPa, respectively.
bricant while tunnelling in gravel formation. The four drives char- The pipe-jacking of Drive B spanned 126 m within the alternating
acteristics are shown in Table 3 (after Cheng et al., 2017a,b). layers of gravel formation and fine soil governed gravel or sand deposit
at a depth of 10.3 m (Fig. 4). Driving into the fine soil governed gravel
3.3. Jacking force baseline establishment or sand deposit at 18–29 m distance led to an increase of 294 kN in the
jacking force and this increase gave away when driving back into the
Figs. 3–6 show the variation in the jacking force and torque of cutter gravel formation at 29 m distance. The jacking force increased again
wheel at the four pipe-jacking drives. Pellet-Beacour and Kastner while driving into the fine soil governed gravel or sand deposit at 45 m
(2002) reported that local variations of total jacking load are generally distance from the gravel formation. The jacking force then decreased
linked to the varying face resistance and that the minima of total when driving back into the gravel formation at 114 m distance.
jacking load correspond to very low face resistances. Thus, a line run- Therefore, there were four sections; the initial 25 m (from 4 to 29 m)
ning through the minima of total jacking load is referred to as “baseline section, 29–45 m section, 45–114 m section, and final 114–123 m sec-
of jacking forces” and can be utilised to distinguish the friction and face tion, included in the baseline of jacking forces. The jacking force in the
resistances from the total jacking load. In this study, each jacking force four baseline sections averaged 0.4, 3.1, 2.1, and 5.4 kN/m, respec-
baseline section was attentively established based upon the minima of tively, inducing the frictional stresses of 0.09, 0.7, 0.5, and 1.2 kPa,

497
W.-C. Cheng et al.

Table 3
Characteristics of the four pipe-jacking drives (after Cheng et al., 2017a,b).

Pipe-jacking drives A B C D

Geology Nearly 90% fine soil governed gravel or 50% gravel formation and 50% fine soil 45% alternating layers of sand and gravel and 55% fine 20% gravel formation and 80% fine soil
sand deposit governed sand deposit soil governed gravel or sand deposit governed gravel or sand deposit
Baseline section Section 1: 1–5 m distance in gravel; Section 1: 4–29 m distance in gravel; Section 1: 2–8 m distance in gravel; Section 1: 6–11 m distance in gravel;
Section 2: 5–67 m distance in fine soil Section 2: 29–45 m distance in gravel; Section 2: 8–21 m distance in gravel; Section 2: 11–24 m distance in fine soil
governed gravel or sand deposit; Section 3: 45–114 m distance in fine soil Section 3: 21–40 m distance in gravel; governed gravel or sand deposit;
Section 3: 67–70 m distance in gravel governed gravel or sand deposit; Section 4: 40–75 m distance in fine soil governed gravel Section 3: 24–31 m distance in fine soil
Section 4: 114–123 m distance in gravel or sand deposit governed gravel or sand deposit;
Section 4: 31–102 m distance in fine soil
governed gravel or sand deposit
Length: m 73 126 75 102

498
Depth: m 10.3 10.3 10.8 10.8
Soil cover to outer pipe diameter 6.7 6.7 7 7
ratio (h/De)
Groundwater level: m 4.5 4.5 4.5 4.5
Jacking alignment (straight or Straight Straight Straight Straight
curve)
Cutter face diameter: m 1.5 1.5 1.5 1.5
Outer pipe diameter De: m 1.44 1.44 1.44 1.44
Measured initial face resistance: kPa 388 499 555 555
Pipe self-weight: kN/m 12.7 12.7 12.7 12.7
Buoyancy force: kN/m 17 17 17 17
Theoretical overcut annulus: litre/m 138 138 138 138
Average volume of lubricant 378 381 552 534
injected: litre/m
Tunnelling and Underground Space Technology 71 (2018) 494–504
W.-C. Cheng et al. Tunnelling and Underground Space Technology 71 (2018) 494–504

2000 200 Fig. 3. Variation of jacking force and torque during tunnelling at drive A
Gravel Clay deposit or Gravel Clay and baseline of jacking forces.
1800 deposit fine soil governed gravel or sand deposit deposit deposit 180

1600 160

1400 140
Jacking force (kN)

favg=1.6 kN/m in
Section 2 (5-67 m) favg=32.7 kN/m in

Torque (A)
1200 favg=36.8 kN/m in 120
Section 3 (67-70 m)
Section 1 (1-5 m)
1000 100

800 80

600 60

400 Average face 40


Variation of torque
resistance
= 685.6 kN (initial contact Jacking force
200 pressure)/area of Baseline of jacking forces 20
cutter wheel = 388 kPa
0 0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70
Jacked distance (m)

2000 200 Fig. 4. Variation of jacking force and torque during tunnelling at drive B
Gravel Gravel Fine soil governed Gravel Gravel and baseline of jacking forces.
1800 deposit deposit sand deposit deposit deposit 180
Fine soil governed Fine soil governed
1600 sand deposit sand deposit 160
favg=2.1 kN/m in
1400 favg=3.1 kN/m in Section 3 (45-114 m) 140
Jacking force (kN)

favg=0.4 kN/m in Section 2 (29-45 m)

Torque (A)
1200 Section 1 (4-29 m) 120
favg=5.4 kN/m in
1000 Section 4 (114-123 m) 100

800 80

600 60
Average face
400 resistance = 40
881.8 kN (initial contact Variation of torque
pressure)/area of Jacking force
200 cutter wheel = 499 kPa Baseline of jacking forces 20

0 0
0 10 20 30 40 50 60 70 80 90 100 110 120
Jacked distance (m)

3000 6 7 9 11 13 200 Fig. 5. Variation of jacking force and torque during tunnelling at drive C
Deviation>60 mm 1 2 3 4 5 Buried wooden log and baseline of jacking forces.
2750 8 10 12 180
1 3 5 (odd number): Gravel deposit
2500 2 4 6 (even number): Fine soil governed gravel or sand deposit
160
2250 favg=0.5 kN/m in
Section 3 (21-40 m) 140
Jacking force (kN)

2000 favg=56.5 kN/m in


Torque (A)

1750 Section 2 (8-21 m) favg=12.3 kN/m in 120


Section 4 (40-75 m)
1500 Variation of torque
100
1250 Jacking force
Baseline of jacking forces 80
1000
60
750
favg=1.6 kN/m in 40
500 Section 1 (2-8 m) Average face
resistance = 980.8 kN (initial contact
250 pressure)/area of 20
cutter wheel = 555 kPa
0 0
0 10 20 30 40 50 60 70
Jacked distance (m)

respectively. tunnelling alignment. The jacking force increase, 771 kN, resulting from
Fig. 5 shows the pipe-jacking activities from Drive C spanning 75 m jacking through the gravel formation between 8 and 21 m distance, was
through a first 34 m gravel section into a 41 m fine soil governed gravel attributed to the excessive pipe deviation greater than a threshold value
or sand deposit at a depth of 10.8 m. Similar to previous two drives, the of 60 mm. The jacking force increased from 1773.8 to 2597 kN between
jacking force increased while driving into the fine soil governed gravel 40 and 51 m distance while driving into the fine soil governed gravel or
or sand deposit and decreased when returning to the gravel formation. sand deposit. Additionally, a buried wooden log encountered when
However, increase in the jacking force can also be caused either by driving in the fine soil governed gravel or sand deposit between 59 and
excessive pipe deviation or by buried obstructions along design 63 m distance geared up the jacking force to an even higher value of

499
W.-C. Cheng et al. Tunnelling and Underground Space Technology 71 (2018) 494–504

4000 200 Fig. 6. Variation of jacking force and torque during tunnelling at drive D
Deviation>60 mm Deviation>60 mm 6 Buried wooden log and baseline of jacking forces.
1 2 3 4 5 7 8 9 10 11 12 13 180
3500
1 3 5 (odd number): Gravel deposit 160
3000 2 4 6 (even number): Silt deposit or fine soil
governed gravel or sand deposit
140
Jacking force (kN)

2500

Torque (A)
favg=49.0 kN/m in 120
Section 3 (24-31 m)
Variation of torque
favg=3.5 kN/m in
2000 favg=18.1 kN/m in
Section 4 (31-102 m)
Jacking force 100
Section 2 (11-24 m) Baseline of jacking forces
80
1500
60
1000
40
Average face
500
favg=2.0 kN/m in resistance = 980.8 kN (initial contact 20
Section 1 (6-11 m) pressure)/area of
cutter wheel = 555 kPa
0 0
0 10 20 30 40 50 60 70 80 90 100
Jacked distance (m)

Table 4 180
Parameters adopted in the jacking force models.
160
Model parameters Value (Maximum) (Minimum)
140
Method 1 (JMTA)
120

pressure [kPa]
Normal contact
ϕ′ (Friction angle): deg. 34 (Gravelly soil)
26.3 (Clayey soil) 100
C′ (Cohesion of the soil): kPa 0
γ′ (Unit weight of the soil): kN/m3 10.09 (Gravelly soil) 80
8.03 (Clayey soil)
De (Outer diameter of the pipe): m 1.44 60
Method 2 (Ma Baosong)
40
Kp (Factor of the vertical soil pressure): kN/m3 0.3 (Gravelly soil)
0.5 (Clayey soil) 20
h (Soil height above the top of pipe): m 9.6 (Drives A & B)
10.1 (Drives C & D) 0
γ′ (Unit weight of the soil): kN/m3 10.09 (Gravelly soil)
8.03 (Clayey soil)
De (Outside diameter of the pipe): m 1.44
ϕ′ (Friction angle): deg. 34 (Gravelly soil)
26.3 (Clayey soil)
ω (Pipe self-weight per meter): kN/m 12.7

Method 3 (Shimada and Matsui)


D1 (Diameter of the bore): m 1.5 Fig. 7. Comparison of the calculated normal contact pressure between the jacking force
D2 (Outer diameter of the pipe): m 1.44 models.
Pu (Pipe self-weight per meter): kN/m 12.7
kd = (D1 × D2)/(D1 − D2) 36
E1 (Young’s modulus of the soil): kPa 22,373 (Gravelly 1300
soil) 400 LEGEND Given parameters:
8829 (Clayey soil) =10 kN/m 3,
JMTA D e=1.44 m,
n1 (Poisson’s ratio of the soil) 0.35 (Gravelly soil) 350 Ma Baosong =30 deg.,
0.49 (Clayey soil) Shimada and Matsui =12.7 kN/m,
300
pressure [kPa]
Normal contact

Terzaghi b=0.214 m
n2 (Poisson’s ratio of the pipe) 0.2
GB 50332
E2 (Young’s modulus of the pipe): kPa 40,000,000 250 ATV A 161
Ce = (1 − n12)/E1 + (1 − n22)/E2: kPa−1 4e−5 (Gravelly soil) ASTM F 1962
9e−5 (Clayey soil) 200 BS EN 1594
Soil prism
b = 1.6 × (Pu × kd × Ce)0.5 = 2 × a (Contact width 0.21 (Gravelly soil)
150
between the pipe and bore): m 0.32 (Clayey soil)
x (Distance to either side of the centerline of the area of 0 100
contact): m
pmax = 2Pu × [1 − (x2/a2)]0.5/(π × a) (Maximum 75.4 (Gravelly soil) 50
normal contact pressure acting on the pipe bottom): 50.9 (Clayey soil)
0
kPa
0 10 20 30 40 50
Method 4 (Pellet-Beaucour and Kastner) h/D e [-]
De (Outer diameter of the pipe): m 1.44
ϕ′ (Friction angle): deg. 34 (Gravelly soil) Fig. 8. Variation of normal contact pressure σ′ against h/De.
26.3 (Clayey soil)
C′ (Cohesion of the soil): kPa 0
γ′ (Unit weight of the soil): kN/m3 10.09 (Gravelly soil) 2793 kN. The jacking force then decreased to 2205 kN at 75 m distance
8.03 (Clayey soil) while re-entering the gravel formation. Thus, the beginning 6 m (from 2
h (Soil height above the top of pipe): m 9.6 (Drives A and B) to 8 m) section, 8–21 m section, 21–40 m section, and final 40–75 m
10.1 (Drives C and
section should be considered in determining the baseline of jacking
D)
forces. The average jacking force in the four baseline sections measured

500
W.-C. Cheng et al. Tunnelling and Underground Space Technology 71 (2018) 494–504

1 the soil arching factor (σ/γh) in Terzaghi and JMTA converged to the
LEGEND
ATV A 161
constant value sooner than the other models (see Fig. 9) and the effect
JMTA
Ma Baosong ASTM F 1962 of soil arching shortly had a steady influence on the σ′ value. It is
Soil arching factor / h [-]

0.8
Terzaghi BS EN 1594 known that the Shimada and Matsui model assumed a stable tunnel
GB 50332
bore and the calculation of σ′ value was not dependent upon the h/De
0.6 value, but upon the values of Pu and b. It can also be seen from
Given parameters: Fig. 8 that as h/De < 7, ASTM F 1962 was the maximum, while JMTA
=10 kN/m3,
De=1.44 m, was the minimum. These results matched with those shown in Fig. 7. As
0.4 =30 deg., ASTM F 1962 made a more conservative assumption that reduces the
=12.7 kN/m
mobilised shearing resistance along shearing bands by taking only half
0.2 the friction angle into account, it thus gave the highest estimation. On
the other hand, JMTA calculated the σ′ value by introducing the re-
duced weight of soil prism that takes the overburden height equal to
0 half the outer pipe diameter into account and considered fully devel-
0 10 20 30 40 50
oped shearing bands. The two assumptions are regarded as the two
h/De [-]
main reasons to lead to the lowest estimation.
Fig. 9. Variation of soil arching factor σ/γh against h/De.
3.5. Back-analysed frictional coefficient
at 1.6, 56.5, 0.5, and 12.3 kN/m, respectively, leading to the frictional
stresses of 0.4, 12.5, 0.1, and 2.7 kPa, respectively. As the jacking force at one end of a baseline section minus that at
The pipe-jacking activities for Drive D spanning 102 m from an in- the other end is equal to the pipe friction along the section, dividing the
itial 17 m section comprising gravel and sand through a following 82 m pipe friction by the section length leads to the average jacking force
alternating layers of clayey gravel and silty sand into a final 3 m gravel (favg). Dividing the favg value by the outer periphery of pipe yields the
section at a depth of 10.8 m are shown in Fig. 6. The jacking force due measured frictional stress (τ). Thus, the frictional coefficient (μavg) can
to the excessive pipe deviation being greater than the threshold value of be back-analysed through dividing the τ value by the σ′ value. The
60 mm increased from 989.8 kN at 11 m distance to 1225 kN at 24 m values of σ′, favg, and τ are summarised in Table 5. In the first and final
distance. A significant increase in the jacking force between 17 and sections at Drive A, the μavg values in JMTA were back-analysed to be
22 m distance was attributed to the varying face resistance resulting 0.59 and 0.52, respectively (Table 6), largely in excess of the upper
from driving into the fine soil governed gravel or sand deposit. The limit of μ = 0.3 suggested by Stein et al. (1989). This indicates that
jacking force declined when re-entering the gravel formation at 31 m lubrication during pipe-jacking in the two sections of gravel at Drive A
distance. The jacking force between 31 and 45 m distance increased to was not effective, despite having nearly 2.7 times more lubricant in-
2842 kN owing to the combined effects of the excessive pipe deviation jected into the theoretical overcut (Table 3). In contrast with JMTA, the
greater than 60 mm and driving into the fine soil governed gravel or other models reported the μavg values below the lower limit of μ = 0.1
sand deposit. Additionally, encountering the buried wooden log be- (Stein et al., 1989). In fact, the higher μavg values of 0.59 and 0.52 were
tween 69 and 76 m distance led the jacking force to increase to a peak caused by excessive pipe frictions incurred while jacking through the
of 3579 kN. The jacking force decreased when returning to the gravel two gravel sections where local collapse or ground closure occurred.
formation at 99 m distance. With the pipe friction incurred between 99 Thus, it can be deduced that JMTA provided the μavg values reflected
and 102 m distance, the jacking force finally slightly increased by what was happened during tunnelling through the sections.
49 kN. Thus, the first 5 m (from 6 to 11 m) section, 11–24 m section, At Drive B, the μavg values from all the models were far below the
24–31 m section, and final 31–102 m section would be included in the lower limit of μ = 0.1, suggesting that Drive B was very well-lubricated.
baseline of jacking forces. The jacking force in the four baseline sections In the second section of gravel at Drive C, the μavg value in JMTA was
averaged 2.0, 18.1, 49.0, and 3.5 kN/m, respectively, inducing the 0.9 (Table 6), largely in excess of the upper limit of μ = 0.3, suggesting
frictional stresses of 0.4, 4.0, 10.8, and 0.8 kPa, respectively. that Drive C through the second section was not adequately lubricated.
In contrast to JMTA, the μavg values from the other models were within
the limits of 0.1 and 0.3 recommended by Stein et al. (1989) for lu-
3.4. Calculated normal contact pressure bricated drives. The μavg value of 0.9, the highest of the μavg values
observed in this study, was most likely due to increasing pipe friction
For the purpose of simplicity, one single stratum has been assumed resulting from the excessive pipe deviation greater than the threshold of
in calculations of the normal contact pressure (σ′). Considering the 60 mm (Fig. 5). The μavg value from JMTA was in line with what was
groundwater table of 4.5 m depth below the ground surface, the σ′ observed while tunnelling through the second section at Drive C, while
value at the four drives can be yielded, respectively, through the con- the other models resulted in a misleading interference for estimation of
tact pressure equation of the four models (see Table 1) by substituting the μavg value.
the parameters listed in Table 4. The maximum and minimum of the σ′ In the second section of silty gravel or silty sand at Drive D, the μavg
value in each model are depicted in Fig. 7, together with the upper value of 0.3 in JMTA matched with the upper limit of μ = 0.3 (Table 6),
bound value resulting from soil prism above the pipe. It can be seen indicating that Drive D through the second section was moderately
from Fig. 7 that the σ′ value varying from 13.8 to 14.8 kPa in JMTA was lubricated. Whereas the other models reported the μavg values below
the minimum amongst all the models, while in ASTM F 1962, the σ′ 0.1. In point of fact, based on the measured data, the μavg value of 0.3
value varying from 101.7 to 122.2 kPa was the maximum. A calculation was attributed to increasing friction resistance resulting from the pipe
sample is given in Fig. 8 for further clarification. It can be seen from deviation being greater than 60 mm (Fig. 6). Similar to the other drives,
Fig. 8 that the σ′ value in ASTM F 1962, ATV A 161, GB 50332, and BS JMTA reported the μavg values more realistic than the other models and
EN 1594 increased slightly at small h/De and maintained at a constant has shown a potential use in assessing the interface behaviour during
value at large h/De. The σ′ value in Ma Baosong increased linearly along pipe-jacking works. Additionally, Drive D through the third section of
with the increasing h/De. The σ′ value in Terzaghi and JMTA as well as silty gravel or clayey gravel characterised by JMTA indicated in-
Shimada and Matsui maintained at constant values of 26.9 and 15.3 as adequate lubrication, with the μavg value of 0.73 (Table 6). The μavg
well as 40.4, respectively, in all the h/De values. The two constant σ′ value of 0.73 was the second highest of the μavg values observed in this
values from Terzaghi and JMTA, respectively, are due to the fact that study. However, the reason to cause this higher μavg value was not due

501
W.-C. Cheng et al. Tunnelling and Underground Space Technology 71 (2018) 494–504

Table 5
Summary of (a) the average jacking forces and measured frictional stresss as well as calculated normal contact pressures based upon (b) JMTA, (c) Ma Baosong, (d) Shimada and Matsui,
(e) Terzaghi, (f) GB 50332, (g) ATV A 161, (h) ASTM F 1962, (i) BS EN 1594.

Pipe-jacking drives A B C D

(a)
Average jacking force favg: kN/m Section 1: 36.8 Section 1:0.4 Section 1: 1.6 Section 1: 2.0
Section 2: 1.6 Section 2:3.1 Section 2: 56.5 Section 2: 18.1
Section 3: 32.7 Section 3:2.1 Section 3: 0.5 Section 3: 49.0
Section 4:5.4 Section 4: 12.3 Section 4: 3.5
Measured frictional stress τ: kPa Section 1: 8.1 Section 1: 0.09 Section 1: 0.4 Section 1: 0.4
Section 2: 0.4 Section 2: 0.7 Section 2: 12.5 Section 2: 4.0
Section 3: 7.2 Section 3: 0.5 Section 3: 0.1 Section 3: 10.8
Section 4: 1.2 Section 4: 2.7 Section 4: 0.8

(b)
Calculated normal contact pressure σ′: kPa Section 1: 13.8 Section 1: 13.8 Section 1: 13.8 Section 1: 13.8
Section 2: 14.8 Section 2: 13.8 Section 2: 13.8 Section 2: 14.8
Section 3: 13.8 Section 3: 14.8 Section 3: 13.8 Section 3: 14.8
Section 4: 13.8 Section 4: 14.8 Section 4: 14.8

(c)
Calculated normal contact pressure σ′: kPa Section 1: 89.8 Section 1: 89.8 Section 1: 92.8 Section 1: 92.8
Section 2: 115.8 Section 2: 89.8 Section 2: 92.8 Section 2: 119.4
Section 3: 89.8 Section 3: 115.8 Section 3: 92.8 Section 3: 119.4
Section 4: 89.8 Section 4: 119.4 Section 4: 119.4

(d)
Calculated normal contact pressure σ′: kPa Section 1: 75.4 Section 1: 75.4 Section 1: 75.4 Section 1: 75.4
Section 2: 50.9 Section 2: 75.4 Section 2: 75.4 Section 2: 50.9
Section 3: 75.4 Section 3: 50.9 Section 3: 75.4 Section 3: 50.9
Section 4: 75.4 Section 4: 50.9 Section 4: 50.9

(e)
Calculated normal contact pressure σ′: kPa Section 1: 58.9 Section 1: 58.9 Section 1: 59.4 Section 1: 59.4
Section 2: 65.3 Section 2: 58.9 Section 2: 59.4 Section 2: 66.1
Section 3: 58.9 Section 3: 65.3 Section 3: 59.4 Section 3: 66.1
Section 4: 58.9 Section 4: 66.1 Section 4: 66.1

(f)
Calculated normal contact pressure σ′: kPa Section 1: 97.7 Section 1: 97.7 Section 1: 99.8 Section 1: 99.8
Section 2: 86.0 Section 2: 97.7 Section 2: 99.8 Section 2: 87.7
Section 3: 97.7 Section 3: 86.0 Section 3: 99.8 Section 3: 87.7
Section 4: 97.7 Section 4: 87.7 Section 4: 87.7

(g)
Calculated normal contact pressure σ′: kPa Section 1: 108.4 Section 1: 108.4 Section 1: 111.1 Section 1: 111.1
Section 2: 99.5 Section 2: 108.4 Section 2: 111.1 Section 2: 102.0
Section 3: 108.4 Section 3: 99.5 Section 3: 111.1 Section 3: 102.0
Section 4: 108.4 Section 4: 102.0 Section 4: 102.0

(h)
Calculated normal contact pressure σ′: kPa Section 1: 118.9 Section 1: 118.9 Section 1: 122.2 Section 1: 122.2
Section 2: 101.7 Section 2: 118.9 Section 2: 122.2 Section 2: 104.3
Section 3: 118.9 Section 3: 101.7 Section 3: 122.2 Section 3: 104.3
Section 4: 118.9 Section 4: 104.3 Section 4: 104.3

(i)
Calculated normal contact pressure σ′: kPa Section 1: 92.4 Section 1: 92.4 Section 1: 94.2 Section 1: 94.2
Section 2: 84.6 Section 2: 92.4 Section 2: 94.2 Section 2: 86.3
Section 3: 92.4 Section 3: 84.6 Section 3: 94.2 Section 3: 86.3
Section 4: 92.4 Section 4: 86.3 Section 4: 86.3

to unfavourable pipe friction resulting from the excessive pipe devia- pressure (σ′) by introducing the reduced overburden height and
tion, but to the gravel formation at 32 m distance not being long enough assuming fully developed shearing bands. As h/De < 7, ASTM F
to establish lower face resistance or total jacking load. 1962 gave the highest estimation as it made a more cautious as-
sumption that reduces the mobilised shearing resistance along the
4. Conclusions sliding planes by taking only half the friction angle into account.
(3) Pipe-jacking data from the four drives were represented using the
From the comparison of the calculated normal contact pressures baseline technique. In JMTA, the back-analysed frictional coeffi-
based upon the four jacking force models and the back-analysed fric- cient (μavg) value, resulting from jacking through the first and final
tional coefficients, the following conclusions can be drawn: sections of gravel at Drive A, was 0.59 and 0.52, respectively, while
the other models reported the μavg values below 0.1. The higher μavg
(1) Four well-established jacking force models were reviewed in this values were most likely due to increasing pipe friction incurred
study. Amongst all the models, the σ/γh value in JMTA is the while jacking through the gravel sections where local collapse or
minimum in all the h/De values, suggesting that the effect of soil ground closure occurred. The μavg values of 0.9 and 0.3, respec-
arching performed best in JMTA. tively, induced by tunnelling through the second section of gravel at
(2) JMTA provided the lowest estimation of the normal contact Drive C and the second section of fine soil governed ground at Drive

502
W.-C. Cheng et al. Tunnelling and Underground Space Technology 71 (2018) 494–504

Table 6
Summary of the back-analysed μavg values for the four pipe-jacking drives.

Pipe-jacking drives A B C D

Geology in baseline section Section 1: Gravel Section 1: Gravel Section 1: Gravel Section 1: Gravel
Section 2: Clayey gravel Section 2: Gravel Section 2: Gravel Section 2: Clayey gravel or silt
or clay Section 3: Clayey sand Section 3: Gravel Section 3: Clayey gravel or silt
Section 3: Gravel Section 4: Gravel Section 4: Clayey gravel Section 4: Clayey gravel or silt
Frictional coefficient μ recommended by Stein et al. (1989) Section 1: 0.3–0.4 Section 1: 0.3–0.4 Section 1: 0.3–0.4 Section 1: 0.3–0.4
Section 2: 0.2–0.3 Section 2: 0.3–0.4 Section 2: 0.3–0.4 Section 2: 0.2–0.3
Section 3: 0.3–0.4 Section 3: 0.2–0.3 Section 3: 0.3–0.4 Section 3: 0.2–0.3
Section 4: 0.3–0.4 Section 4: 0.2–0.3 Section 4: 0.2–0.3
Back-analysed μavg value by JMTA (2000) Section 1: 0.59 Section 1: 0.007 Section 1: 0.03 Section 1: 0.03
Section 2: 0.03 Section 2: 0.05 Section 2: 0.9 Section 2: 0.30
Section 3: 0.52 Section 3: 0.03 Section 3: 0.007 Section 3: 0.73
Section 4: 0.09 Section 4: 0.18 Section 4: 0.05
Back-analysed μavg value by Ma (2008) Section 1: 0.09 Section 1: 0.001 Section 1: 0.004 Section 1: 0.004
Section 2: 0.003 Section 2: 0.008 Section 2: 0.14 Section 2: 0.03
Section 3: 0.08 Section 3: 0.004 Section 3: 0.001 Section 3: 0.09
Section 4: 0.01 Section 4: 0.02 Section 4: 0.007
Back-analysed μavg value by Shimada and Matsui (1998) Section 1: 0.11 Section 1: 0.001 Section 1: 0.005 Section 1: 0.005
Section 2: 0.008 Section 2: 0.009 Section 2: 0.17 Section 2: 0.08
Section 3: 0.10 Section 3: 0.01 Section 3: 0.001 Section 3: 0.21
Section 4: 0.02 Section 4: 0.05 Section 4: 0.02
Back-analysed μavg value by Terzaghi (1951) Section 1: 0.14 Section 1: 0.002 Section 1: 0.007 Section 1: 0.007
Section 2: 0.006 Section 2: 0.01 Section 2: 0.21 Section 2: 0.06
Section 3: 0.12 Section 3: 0.008 Section 3: 0.002 Section 3: 0.16
Section 4: 0.02 Section 4: 0.04 Section 4: 0.01
Back-analysed μavg value by GB 50332-02 (2002) Section 1: 0.08 Section 1: 0.001 Section 1: 0.004 Section 1: 0.004
Section 2: 0.005 Section 2: 0.007 Section 2: 0.13 Section 2: 0.05
Section 3: 0.07 Section 3: 0.006 Section 3: 0.001 Section 3: 0.12
Section 4: 0.01 Section 4: 0.03 Section 4: 0.009
Back-analysed μavg value by ATV A 161 (German ATV rules Section 1: 0.08 Section 1: 0.001 Section 1: 0.004 Section 1: 0.004
and standards, 1990) Section 2: 0.004 Section 2: 0.006 Section 2: 0.11 Section 2: 0.04
Section 3: 0.07 Section 3: 0.005 Section 3: 0.001 Section 3: 0.11
Section 4: 0.01 Section 4: 0.03 Section 4: 0.008
Back-analysed μavg value by ASTM F 1962 (ASTM, 2011) Section 1: 0.07 Section 1: 0.001 Section 1: 0.003 Section 1: 0.003
Section 2: 0.004 Section 2: 0.006 Section 2: 0.10 Section 2: 0.04
Section 3: 0.06 Section 3: 0.005 Section 3: 0.001 Section 3: 0.10
Section 4: 0.01 Section 4: 0.03 Section 4: 0.008
Back-analysed μavg value by BS EN 1594 (BS, 2009) Section 1: 0.09 Section 1: 0.001 Section 1: 0.004 Section 1: 0.004
Section 2: 0.005 Section 2: 0.008 Section 2: 0.13 Section 2: 0.05
Section 3: 0.08 Section 3: 0.006 Section 3: 0.001 Section 3: 0.13
Section 4: 0.01 Section 4: 0.03 Section 4: 0.009
Tunnel bore condition deduced by JMTA Section 1: Ground Section 1: Stable tunnel Section 1: Stable tunnel Section 1: Stable tunnel bore
closure bore bore Section 2: Possible ground
Section 2: Stable tunnel Section 2: Stable tunnel Section 2: Possible ground closure
bore bore closure Section 3: Not applicable
Section 3: Ground Section 3: Stable tunnel Section 3: Stable tunnel Section 4: Stable tunnel bore
closure bore bore
Section 4: Stable tunnel Section 4: Ground closure
bore

D were attributed to increasing friction resistance resulting from the valuable comments from two anonymous reviewers are much appre-
excessive pipe deviation greater than the threshold value of 60 mm. ciated.
While the other models reported the μavg values within the limits of
0.1 and 0.3 and below 0.1, respectively. References
(4) Pipe-jacking through the third section of silty gravel or clayey
gravel at Drive D characterised by JMTA indicated inadequate lu- ASTM, 2002. Standard Test Method for Particle-Size Analysis of Soils. D422-63, West
brication, with the μavg value of 0.73. However, the higher μavg Conshohocken, PA.
ASTM, 2011. Standard guide for use of maxi-horizontal directional drilling for placement
value of 0.73 was not caused by unfavourable pipe friction resulting of polyethylene pipe or conduit under obstacles including river crossings. F 1962-11,
from the excessive pipe deviation, but by the gravel formation at West Conshohocken, PA.
32 m distance not being long enough to establish lower face re- Barla, M., Camusso, M., Aiassa, S., 2006. Analysis of jacking forces during microtunnel-
ling in limestone. Tunn. Undergr. Space Technol. 21 (6), 668–683. http://dx.doi.org/
sistance or total jacking load. 10.1016/j.tust.2006.01.002.
(5) JMTA yielded the μavg value more realistic than the other models British standards, 2009. BS EN:1594-09 Gas supply system-pipelines for maximum op-
and has shown a potential use in assessing the interface perfor- erating pressure over 16 bar-functional requirements. Brussels, pp. 76–78.
Chapman, D.N., Ichioka, Y., 1999. Prediction of jacking forces for microtunnelling op-
mance during pipe-jacking works.
erations. Tunn. Undergr. Space Technol. 14 (1), 31–41. http://dx.doi.org/10.1016/
S0886-7798(99)00019-X.
Acknowledgements Choo, C.S., Ong, D.E.L., 2015. Evaluation of pipe-jacking forces based on direct shear
testing of reconstituted tunneling rock spoils. J. Geotech. Geoenviron. Eng. 141 (10),
04015044. http://dx.doi.org/10.1061/(ASCE)GT.1943-5606.0001348.
This study would not have been possible without the financial and Cui, Q.L., Xu, Y.S., Shen, S.L., Yin, Z.Y., Horpibulsuk, S., 2015. Field performance of
technical support from the Kao Kun Construction Co. Ltd. A series of concrete pipes during jacking in cemented sandy silt. Tunn. Undergr. Space Technol.

503
W.-C. Cheng et al. Tunnelling and Underground Space Technology 71 (2018) 494–504

49, 336–344. http://dx.doi.org/10.1016/j.tust.2015.05.005. with tidal, partial penetration, and storage effects. Soils Found. 53 (6), 894–902.
Cheng, W.C., Ni, J.C., Shen, S.L., Huang, H.W., 2017a. Investigation into factors affecting Ni, J.C., Cheng, W.C., 2014. Quality control of double fluid jet grouting below ground-
jacking force: a case study. P. I. Civ. Eng. Geotec. 170 (4), 322–334. http://dx.doi. water table: case history. Soils Found. 54 (6), 1039–1053.
org/10.1680/jgeen.16.00117. Namli, M., Guler, E., 2017. Effect of bentonite slurry pressure on interface friction of pipe
Cheng, W.C., Ni, J.C., Shen, S.L., 2017b. Experimental and analytical modeling of shield jacking. J. Pipeline Syst. Eng. Pract. 8 (2), 04016016.
segment under cyclic loading. Int. J. Geomech. 17 (6), 04016146. http://dx.doi.org/ Pellet-Beacour, A.L., Kastner, R., 2002. Experimental and analytical study of friction
10.1061/(ASCE)GM.1943-5622.0000810. forces during microtunneling operations. Tunn. Undergr. Space Technol. 17 (1),
Cheng, W.C., Ni, J.C., Cheng, Y.H., 2017c. Alternative shoring for mitigation of pier- 83–97.
foundation excavation disturbance to an existing freeway. J. Perform. Constr. Fac. 31 Roark, R.J., Young, W.C., 1976. Formulas for Stress and Strain. McGraw-Hill, New York.
(5), 04017072. http://dx.doi.org/10.1061/(ASCE)CF.1943-5509.0001063. Rogers, C.D.F., Yonan, S.J., 1992. Experimental study of a jacked pipeline in sand.
D'Appolonia, D.J., 1980. Soil-bentonite slurry trench cutoffs. J. Geotech. Eng. 106 (4), Tunnels Tunn. Int. 24 (6), 35–38.
399–417. Rahjoo, S., Najaf, M., Williammee, R., Khankarli, G., 2012. Comparison of jacking load
French Society for Trenchless Technology, 2006. Microtunnelling and Horizontal Drilling: models for trenchless pipe jacking. Pipelines 2012 – Innovations in Design,
French National Project “Microtunnels”: Recommendations. ISTE, London. Construction, Operations, and Maintenance, Florida, pp. 1507–1520.
German ATV rules and standards, 1990. ATV-A 161 E-90. Structural Calculation of Driven Stein, D., Möllers, K., Bielecki, R., 1989. Microtunneling: Installation and Renewal of
Pipes. Hennef, pp. 18–20. Nonman-Size Supply and Sewage Lines by the Trenchless Construction Method.
Japan Microtunnelling Association (JMTA), 2000. Pipe-jacking Application, Japan Ernst, Berlin, Germany.
Microtunnelling Association (JMTA), Tokyo. Shimada, H., Matsui, K., 1998. A new method for the construction of small diameter
Kastner, R., Pellet-Beacour, A.L., Ouvry, J.F., Guilloux, A., 1996. In-situ monitoring of tunnels using pipe jacking. Proc. of Regional Symposium on Sedimentary Rock
microtunneling projects. Proceedings of International NO-DIG'96, New Orleans, pp. Engineering, pp. 234–239.
171–182. Sofianos, A.I., Loukas, P., Chantzakos, C., 2004. Pipe jacking a sewer under Athens. Tunn.
Khazaei, S., Shimada, H., Matsui, K., 2004. Analysis and prediction of thrust in using Undergr. Space Technol. 19 (2), 193–203. http://dx.doi.org/10.1016/S0886-
slurry pipe jacking method. Proceedings of the 30th ITA-AITES World Tunnel 7798(03)00108-1.
Congress, Singapore, pp. 22–27. Staheli, K., 2006. Jacking Force Prediction an Interface Friction Approach Based on Pipe
Kou, H.L., Chu, J., Guo, W., Zhang, M.Y., 2015. Field study of residual forces developed in Surface Roughness. Ph.D. Thesis, Georgia Institute of Technology.
pre-stressed high-strength concrete (PHC) pipe piles. Can. Geotech. J. 53 (4), Shou, K.J., Yen, J., Liu, M., 2010. On the frictional property of lubricants and its impact
696–707. on jacking force and soil–pipe interaction of pipe-jacking. Tunn. Undergr. Space
Milligan, G.W.E., Norris, P., 1996. Site-based research in pipe jacking – objectives, pro- Technol. 25 (4), 469–477. http://dx.doi.org/10.1016/j.tust.2010.02.009.
cedures and a case history. Trenchless Technol. Res. 11 (1), 3–24. Shen, S.L., Cui, Q.L., Ho, C.E., Xu, Y.S., 2016. Ground response to multiple parallel mi-
Ma, B., 2008. The Science of Trenchless Engineering. China Communications Press, crotunneling operations in cemented silty clay and sand. J. Geotech. Geoenviron.
Beijing. Eng. 142 (5), 04016001.
Norris, P., 1992. The Behavior of Jacked Concrete Pipes During Site Installation. Ph.D. Shen, S.L., Wang, Z.F., Cheng, W.C., 2017. Estimation of lateral displacement induced by
Thesis, University of Oxford. jet grouting in clayey soils. Geotechnique 67 (7), 621–630. http://dx.doi.org/10.
Ni, J.C., Cheng, W.C., Ge, L., 2011. A case history of field pumping tests in a deep gravel 1680/jgeot.16.P.159.
formation in the Taipei Basin, Taiwan. Eng. Geol. 117 (1–2), 17–28. Terzaghi, K., 1936. The shearing resistance of saturated soils and the angle between the
Ni, J.C., Cheng, W.C., 2011a. Shield machine disassembly in grouted soils outside the planes of shear. Proc. of the 1st International Conference on Soil Mechanics and
ventilation shaft: a case history in Taipei Rapid Transit System (TRTS). Tunn. Foundation Engineering. Harvard University Press, Cambridge, MA, pp. 54–56.
Undergr. Space Technol. 26 (2), 435–443. Terzaghi, K., 1943. Theoretical Soil Mechanics. John Wiley & Sons Inc, New York, USA,
Ni, J.C., Cheng, W.C., 2011b. Steering efficiency of microtunnelling in various deposits. pp. 66–76.
Proceedings of the GeoHunan International Conference 2011, Geotech. Special Publ. Terzaghi, K., 1951. Theoretical Soil Mechanics. J. Wiley and Sons, New York.
No. 221. Geo-Inst. ASCE Publs., pp. 33–39. The Ministry of Construction of China, 2002. GB 50332-02, Structural Design Code for
Ni, J.C., Cheng, W.C., 2012a. Steering characteristics of microtunnelling in various soils. Pipeline of Water Supply and Waste Water Engineering. Beijing, pp. 11–12.
Tunn. Undergr. Space Technol. 28, 321–330. http://dx.doi.org/10.1016/j.tust.2011. Wham, B.P., Argyrou, C., O’Rourke, T.D., 2016. Jointed pipeline response to tunneling-
11.003. induced ground deformation. Can. Geotech. J. 53 (11), 1794–1806.
Ni, J.C., Cheng, W.C., 2012b. Characterising the failure pattern of a station box of Taipei Yen, J., Shou, K., 2015. Numerical simulation for the estimation the jacking force of pipe
Rapid Transit System (TRTS) and its rehabilitation. Tunn. Undergr. Space Technol. jacking. Tunn. Undergr. Space Technol. 49, 218–229. http://dx.doi.org/10.1016/j.
32, 260–272. tust.2015.04.018.
Ni, J.C., Cheng, W.C., Ge, L., 2013. A simple data reduction method for pumping tests

504

You might also like