Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

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

net/publication/263426180

Installation of Offshore Driven Piles - Regional Experience

Conference Paper · August 2012


DOI: 10.3850/978-981-07-1896-1

CITATION READS

1 6,788

2 authors, including:

Kaushik Mukherjee
PETRONAS
9 PUBLICATIONS 58 CITATIONS

SEE PROFILE

All content following this page was uploaded by Kaushik Mukherjee on 27 June 2014.

The user has requested enhancement of the downloaded file.


July 24, 2012 16:23: RPS: RPS-MOSS 2012

INSTALLATION OF OFFSHORE DRIVEN PILES –


REGIONAL EXPERIENCE

K. Mukherjeea and N. Nagarajub

McDermott Asia Pacific Pte. Ltd., Singapore.


E-mail: a kaushik.mj@petronas.com.my, b nnagaraju@mcdermott.com

The types of subsoil dealt with in the Gulf of Thailand and surrounding region are
conducive to the installation of driven piles. Among the various aspects of driven
pile design, pile driveability analysis plays a major role in the selection of the fea-
sible hammer and determining pile make-up to suit, besides defining acceptance
criteria to ensure a certain level of confidence of its fitness for purpose. The accu-
racy of a driveability study particularly depends on the selection of its parameters
to best represent the in-situ scenario. The most uncertain of the parameters are
those related to the soil model. It has been observed in recent projects in the Gulf of
Thailand, that the consideration of suitable soil resistance to driving, coupled with
soil dynamic parameters, significantly affects driving sequence planning, design,
material, hammer selection and above all pile acceptance criteria. This is particu-
larly true for a region that contains predominantly clayey subsoil resulting in large
diameter, long friction piles susceptible to driving disturbances and slow strength
recovery. Different researchers suggested suitable dynamic parameters and soil
resistance during driving calibrated to specific regions. Our regional experience
suggests a set of soil degradation and parameters to suit the different stages of
driving which is corroborated by pile monitoring records in one of the sites dur-
ing continuous and delayed driving scenarios. The most sensitive of all dynamic
soil parameters is the skin damping in clay. While Smith’s wave theory approach
considers a conservative assumption in selecting the parameters in absence of
site-specific data, a different approach may be adopted based on consistency of
the clay for a site where the geotechnical parameters are well defined. This paper
includes an independent study on the effects of variations in dynamic parameters
and resistance at different stages of driving and suitability of choices of different
pile make-up and hammers for an optimum design. Earlier experiences, simulation
studies and back-analyses of available driving records are undertaken in order to
reach a reasonable judgment for the region.
Keywords: Friction piles, Driveability, Clay, Dynamic parameters, Skin damping.

1. INTRODUCTION

Fixed platforms are generally supported in the in-place condition on a foundation of driven
piles or drilled and grouted piles depending on the subsoil condition. The piles may be
driven/drilled inside the jacket legs or through individual skirt legs or grouped in clusters

2nd Marine Operations Specialty Symposium (MOSS 2012)


Edited by Yoo Sang Choo, David N. Edelson and Trevor Mills
Copyright © 2012 by MOSS 2012 Organizers. All rights reserved.
ISBN: 978-981-07-1896-1 :: doi:10.3850/978-981-07-1896-1 MOSS-07 271
July 24, 2012 16:23: RPS: RPS-MOSS 2012

272 2nd Marine Operations Specialty Symposium

at the base of each leg. The types of subsoil dealt with in the Gulf of Thailand and surround-
ing region are conducive to the installation of driven piles only. The design of an offshore
pile is fundamentally the optimization of pile length, size and make-up of those conditions
which are suitable to the requirement and supply. An interactive effort combining aspects
related to geotechnical, structural and constructability are required to converge into a fea-
sible, safe and economic solution. Generally the requirement is defined by the purpose and
environment. Other than the checking of the installation feasibility using a set of hammers
in a given environment, driveability analysis is also used to define acceptance criteria for
a majority of the offshore piles. This is particularly important for clayey subsoil where
the piles are long and friction dominated. Unlike short and end-bearing piles, these piles
often cannot be rested on dense strata where they can be met by a definite set of criteria.
Usually a measurement of blow counts, determined and defined by driveability analyses,
forms the acceptance criteria to ensure a certain level of confidence of fitness for the pur-
pose, i.e. development of ultimate capacity for which the pile is designed with optimum
penetration. While uncertainty of the soil strata at target penetration may result in possi-
ble overdrive, an unexpected delay, or occurrence of an unidentified cemented layer, may
cause early refusal or an under-drive scenario. Accuracy of the driveabilty study as well
as the precautionary measures to avoid or minimize delays during driving are of equal
importance.
The accuracy of a driveability study particularly depends on the selection of its parame-
ters which best represent the in-situ scenario. The parameters related to piles are primarily
defined by the ultimate design requirement, whereas those related to selected hammers
are generally supplied by the manufacturer. Any alteration to these parameters fine-tuned
from previous experience or measurements can easily be captured during analyses. The
most uncertain of the parameters are those related to the soil model [static & dynamic].
This paper focuses on an independent study of the effects of variations in soil model
parameters and resistance at different stages of driving and suitability of choices of dif-
ferent pile make-up and hammers for an optimum design in the typical Gulf of Thailand
subsoil strata. The paper takes into account previous experience, simulation studies and
back-analyses of available driving records.

2. OVERVIEW

The complicated system of hammer-pile-soil interaction and its solution through wave
equation analogy forms the basis of driveablity analysis. The strength parameters of the
subsoil are used to develop the soil resistance to driving [SRD] in much the same way as
it is done to calculate the static capacity. Apart from the method suggested by ISO 19902
[2007] or API RP 2A [2000], the methods suggested by several other authors, e.g., Toolan
and Fox [1977], Alm and Hamre [1998], Stevens [1982] etc. are also used to develop SRD
for soils in various geographical regions. In a dynamically driven scenario, the capacity
developed for the static condition is suitably degraded depending on the soil type and
condition. When a pile is driven in clay, there is a temporary reduction in the shear strength
partly due to remoulding and development of undrained conditions within the smear-
or slip-zone surrounding the pile. Phenomenon like ‘friction fatigue’ is also responsible
for a reduced shear strength of a long pile, but is normally treated as a long-term effect,
particularly in clayey subsoil, and is generally taken care of in the ultimate capacity design.
July 24, 2012 16:23: RPS: RPS-MOSS 2012

Installation of Offshore Driven Piles – Regional Experience 273

Some of the methods developed later to calculate SRD in cohesionless strata take this into
account.
Dynamic soil parameters are needed to define soil viscous-damping effects and
elastic deformation or quake. Roussel [1979] and many other researchers indicate that
damping parameters vary widely with the consistency of the clay. Since the friction piles
are usually longer for a site that predominantly consists of clayey subsoil, it is obvious
that the driveability analyses would be sensitive to those parameters. Software such as
GRLWEAP, recommends that the resistance, due to viscous-damping, is calculated accord-
ing to Smith’s original approach, and includes a few proposed damping models similar
to those originally proposed by Smith. However, since there are no direct links between
Smith’s model and standard geotechnical soil test parameters, several investigators of the
dynamic behavior of piling have expressed concern that the current approach is unreliable
for either previously untested soil conditions or certain extreme conditions (e.g., very high
or very low velocities) for which no experience base exists (Rausche, Goble, Likins, 1992).

3. THEORETICAL APPROACH

The theoretical approach generally adopted for the pile driveability analyses has been
briefly enumerated in the following sub-sections.

3.1. Pile Driving Formulation

From earlier days, pile driving formulae gained popularity for their simplicity, and are
still widely used for driving piles to the design capacity. Because of the enormous efforts
by many researchers to develop such formulae, a large number are available in practice.
Among them are popular ones such as Janbu, Hiley or Engineering News Record [ENR]
which are used worldwide. Most formulae are site specific and have their own limita-
tions or uncertainties which need to be offset by including an appropriate factor for safety.
Smith’s [1960] approach of one dimensional wave propagation theory helps us to under-
stand the complicated driveability phenomenon, and was probably the first rationalized
solution to the problem. With the advent of microcomputers, this approach gained substan-
tial popularity over the time. Similar wave propagation approaches were also adopted by
researchers to develop pile driving formulae to manage pile driving in specific cases [Uto
et. al., 1992; Weele and Schellingerhout, 1994; Chen et. al., 1997]. Programs like GRLWEAP
based on Smith’s approach may also be used judiciously to calibrate the assumptions and
safety factors of a suitable pile driving formula for its site-specific usage.

3.2. Initial Impact and Hammer Selection

The hammer-pile-soil interaction can be depicted by a simple model as shown in Figure 1.


The interaction problem may be separated into two steps: (1) the interaction between pile
and hammer and (2) the penetration of the pile into the ground due to the impact generated
by the hammer or dynamic pile-soil interaction. It is difficult to set up an accurate model
for the interaction of pile and hammer because of the complexity of the driving equipment.
Chen et. al. [2003] explored the fundamental solution provided by Clough and Penzien
(1975) to solve the primary problem of hammer impact [Step (1)]. The pile driving system is
July 24, 2012 16:23: RPS: RPS-MOSS 2012

274 2nd Marine Operations Specialty Symposium

Impact Velocity, V h

1
Hammer Mass, m
Pile Cushion
Stifness, k

Shaft Resistance
of Soil, Rs
2
Pile Stiffness, Km
Soil Stiffness
Soil Damping at
at Shaft, K'm
Pile Shaft, Js

Static Point
Resistance, Rt
Soil Stiffness at Soil Damping at
Pile Toe, K'p Pile Toe, Jp

Figure 1. Pile Driving Model.

modeled by three basic elements: a relatively rigid ram, a linear spring (pile anvil/cushion)
and an elastic pile. The solution yields the impact force induced at the pile top during the
first wave passage, as expressed by the Eq. (1).

 
k · Vh −ξ ·ω ·t
Ft = .( e ) . sin (ωd · t) for 0 ≤ ωd · t ≤ π (1)
ωd
where,
 
k k1 ·c
ω := , ωd := ω · 1 − ξ 2 and ξ :=
m 2·ω·E·A
‘k’ is the combined stiffness of the hammer and pile top cushion/anvil assembly, ‘m’ is the
hammer mass, ‘ω’ is the system natural frequency, ‘ωd ’ is the damped frequency, ’ξ’ is the
damping ratio of the system, ‘E’ and ‘A’ are the elastic modulus and cross-sectional area of
the pile material and ‘c’ is the wave velocity.
‘Vh ’ is ram impact velocity, obtained from the kinetic hammer energy (Et ) generally
mentioned by hammer manufacturers [usually maintained around 5–6 m/s for hydraulic
hammers],

2 · e f · Et
Vh :=
m
where ‘e f ’ is the hammer efficiency that describes the percentage of maximum hammer
energy transferred to the pile and soil or the ratio of the kinetic energy available to do work
on the driving system. This depends on various factors, e.g., losses due to friction, mis-
alignment, low stroke, pre-ignition etc. which are different for different types of hammer.
July 24, 2012 16:23: RPS: RPS-MOSS 2012

Installation of Offshore Driven Piles – Regional Experience 275

Hammer Initial Impact Effect of Assembly Stiffness


Clough & Penzien [1975] Solution on Initial Impact

140000

150000
120000

100000
100000

50000

F(t), kN
80000
F(t), kN

0 60000

0.0 5.0 10.0 15.0 20.0


-50000 40000

20000
-100000
0
-150000 0.0 2.0 4.0 6.0 8.0 10.0

t, ms t, ms

(a) (d)

Simplified Initial Impact Effect of Dropheight


on Initial Impact [Drop Weight constant]
of the Hammer Blow
140000
140000

120000 120000

100000 100000
F(t), kN

80000

F(t), kN
80000

60000 F0
60000

40000
40000

20000 t1
t0 20000

0
0
0.0 1.0 2.0 3.0 4.0 5.0 6.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0
t, ms t, ms

(b) (e)
Effect of Hammer on Initial Impact Effect of Drop Weight
on Initial Impact [Dropheight constant]
180000
200000
160000

160000 140000

MHU800 120000
F(t), kN

120000
MHU1200
F(t), kN

100000

MHU1700 80000
80000
60000

40000 40000

20000
0
0
0.0 2.0 4.0 6.0 8.0 0.0 2.0 4.0 6.0 8.0

t, ms
t, ms

(c) (f)

Figure 2. Initial Impact Force.

Usually, for hydraulic hammers the maximum operating efficiency may reach as high as
95% whereas for the steam hammers this can reach up to a maximum of 67–70%. In Smith’s
original model, the solution iteration starts from this ram impact velocity and impact forces
are calculated numerically in each time step.
The solution plot of Eq. (1) takes the shape of a damped sinusoid that shows the variation
of impact force with time [Figure 2(a)]. Figure 2(b) shows the truncated first half plot of
the sinusoid that shows the generated impact on the pile. This plot is characterized by
three distinct parameters describing the shape: the peak impact force ‘Fo ’, duration ‘to ’ and
loading period ‘t1 ’. These are defined in Figure 2(b):
  
k · Vh  −ξ ·ω ·t1  π 1 ω·d
Fo := . e . sin (ωd · t1 ) to := t.1 := . tan−1
ωd ωd ω·d ω·ξ
Hammer Selection: This analytical calculation is particularly important for the primary selec-
tion of the hammer. The drop weight and height has definitive influence on the impact peak
July 24, 2012 16:23: RPS: RPS-MOSS 2012

276 2nd Marine Operations Specialty Symposium

force (Fo ) and impact duration (to ) which in turn influences the penetration efficiency of
the hammer pile system. There is a general notion that lightweight piles could be driven
faster with more efficiency but this is only true for shorter piles with light loading. For
longer piles driven through stiffer media, pile stiffness plays an important role in carrying
large forces down into the ground. The following rules of thumb may be exercised while
selecting a hammer:

• The greater the mass of the hammer, the larger the impact peak force and the longer
the impact duration [Figures 2(c) & 2(f)]
• A greater drop height, keeping the same hammer mass, may also produce a higher
peak force but may not increase the impact duration [Figure 2(e)]
• A greater stiffness of the hammer cushion assembly could also increase the peak
force but would vary the impact duration [Figure 2(d)]
• The optimum hammer should have higher peak force and longer impact dura-
tion to overcome soil resistance and achieve deeper penetration, however, heavier
impact should not destroy/overstress the pile top
• It is advisable to use a larger hammer with low drop height rather than a smaller
hammer with higher drop height
• Because of continuous blow, the cushion gets stiffer as the driving progresses
and this rapidly increases the peak force and decreases the impact duration
[Figure 2(d)]. The train of excessive peak force reduces driving efficiency and tends
to damage the pile top. The optimal stiffness of pile cushion is essential for efficient
driving.

3.3. Driveability and Soil Mechanics

The second step of the driveability analysis is the penetration of the pile in the ground
with the effect of initial hammer impact. In Smith’s driveability model, the analysis is an
integrated approach that consists of time-step solution of a one dimensional wave equa-
tion until the generated wave velocity is fully expended and/or displacement conver-
gence is achieved. This step is indicated in Figure 1. The soil resistance is modeled with
both dynamic and static components. The idealistic static soil model devised by Smith is
depicted in Figure 3. While the static resistance is displacement dependent, and is defined
by an ultimate strength [Ru ] and a “quake” value [Q], the dynamic resistance depends
on the relative motion between the pile and the soil in the slip-zone [shaft friction] and
pile tip.
The quake represents elastic or temporary soil compression, and is inversely related to
soil stiffness. Typical quake values are in the 2 to 5 mm range; soils with high rebound
have considerably higher end bearing quake values, sometimes in excess of 25 mm. Ideal-
istically, the soil would enter the plastic zone beyond this quake displacement and would
undergo plastic displacement or permanent set ‘s’ as indicated in Figure 3, before entering
into the elastic rebound [equal to Q] or unloading stage.
The speed of pile penetration into the ground depends on pile soil interaction. Some
piles penetrate faster than others in the same soil strata while the same pile would behave
differently in different subsoil. Ground offers more resistance to rapidly moving piles than
slower moving ones. This necessitates introducing velocity dependent viscous damping
July 24, 2012 16:23: RPS: RPS-MOSS 2012

Installation of Offshore Driven Piles – Regional Experience 277

Q Soil under
elastic
s
compression
Till the end
of plastic
movement

Probable Impact of
Stress

Q Soil Stiffness

Probable Impact of
viscous damping

Ru
s Final Soil
Original compression or
Soil permanent set
Model after elastic
rebound

s Q
Strain

Figure 3. Smith’s Soil Model.

parameters [dashpot in Figure 1]. The probable impact of viscous damping effect on the
soil resistance model is shown in Figure 3 by dotted lines. Because of dynamic movement
this resistance is instantaneous in nature and does not affect the static capacity of the pile
as a whole. The soil damping parameters affect both the frictional and toe movement of the
pile and at times have substantial effects on overall driveability results. These are discussed
in more detail in the following section.

3.4. Soil Resistance to Driving [SRD]

The methods commonly used to determine resistance to driving are meant for North Sea
sands and clays. These methods are generally extended for use in other regions
worldwide. Sometimes methods those are meant for axial capacity design, e.g., API/ISO or
UWA method, are modified taking into account of consolidation, equalization and ageing
or time effects results from changes in total stress and pore pressure due to soil displace-
ment during installation [cavity expansion] to calculate SRDs. For Gulf of Thailand [GoT]
clay, generally the API-based method is the most common. Some of the common methods
available for calculating SRDs include:

• API/ISO Method
• Toolan and Fox Method
• Steven’s Method
• Semple & Gemeinhardt
• Alm and Hamre
• UWA-05 Method
• Heerema [Proprietary Methods]

A range of resistance is usually determined to provide upper and lower bound to the
expected response of the soils. An upper bound profile would be expected to provide
July 24, 2012 16:23: RPS: RPS-MOSS 2012

278 2nd Marine Operations Specialty Symposium

the most pessimistic hammer blow count behavior. If the measured response is less than
the lower bound limit, then the assumptions made to determine the soil parameters might
need to be examined and pile capacity, which also depends on the assumed soil parame-
ters, might then be modified [typical CAPWAP type of analysis].
The behavior of the pile during driving may result in a plugged or un-plugged response.
The size of the pile may have a significant effect on the response - small diameter pipes,
such as conductors, may respond in a plugged response, large piles will behave as
unplugged (cored). Often, the minimum of the calculated response is assumed, as in the
Toolan and Fox method. However, some methods, e.g. Stevens, recommend the use of
plugged behavior based on back-analysis.
The Toolan and Fox (1977) method has been calibrated against North Sea driving condi-
tions and is generally considered to give conservative i.e., high SRD values. This, in turn,
leads to higher damage and a conservative estimate of fatigue lives. In the Toolan and
Fox method, driving resistance is determined using an upper and lower bound resistance.
In sands measured qc values are used to provide end bearing (base area/annulus area
×0.6 qc ) and skin friction (qc /300) using the limits given in API (for skin friction). The
design SRD value is taken as the minimum of plugged and coring responses. Remoulded
shear strength, calculated using published correlations between liquidity index and
remoulded strength, is used to determine resistance in clays. For a lower bound resis-
tance, the tip resistance may be calculated using say 0.3 qc and the shaft friction may be
determined using half the upper bound value i.e. qc /600.
The Alm and Hamre (2001) method was calibrated primarily based on driveability stud-
ies of large diameter piles in North Sea. For sands, pile diameters ranged from 2.4 m to
2.7 m, and pile penetrations ranged from 55 m to 70 m. To minimize additional uncertain-
ties related to selection of soil parameters and modeling changes in radial stress due to pile
installation, both base resistance (qb ) and skin friction (τ f ) are correlated to CPT ‘qt ’. Shaft
friction is estimated to exponentially decay from a maximum [τf,max ] to a residual value
[τf,res ].
Unlike the Alm and Hamre method, the UWA-05 (Lehane et. al., 2005, 2007) was devel-
oped to estimate static axial capacity in uncemented siliceous sand, rather than SRD. Shaft
friction is evaluated using the Coulomb failure criteria as a function of radial effective
stress, σrf during installation conditions. The radial stress after installation [σrc
 ], consider-

ing the friction fatigue effect and the change in radial stress [Δσrd ], are estimated separately,
and both typically use correlations to cone resistance, qt . The change in radial stress during
loading is based on elastic cylindrical cavity expansion theory (Lehane et. al., 1993).
API RP 2A or ISO19902 method, like UWA-05, is also principally developed to estimate
static axial capacity rather than SRD. With the applied variation for remoulded skin friction
in clay during driving, this method predicts SRD in marine clays reasonably well and is
popularly used for long friction piles in similar conditions to the Gulf of Thailand.

3.5. Dynamic Soil Model

The Smith’s soil model as depicted in Figure 3 does not include the effect of time, i.e. speed
of pile penetration into the ground and the relative velocity of the pile with respect to the
surrounding soil. Smith dealt with this problem by introducing a damping factor ‘J’ to
cater for this viscous damping effect. Thus the part of the actual energy transmitted to the
July 24, 2012 16:23: RPS: RPS-MOSS 2012

Installation of Offshore Driven Piles – Regional Experience 279

soil caters for the losses incurred by the dashpot. Idealistically, the total resistance offered
by the soil would be

Rt = Rs [1 + J.v] (2)

where Rt is the total resistance, Rs is the static part of the resistance, ‘v’ is the relative
velocity of the pile and ‘J’ is the Smith’s damping parameter. The parameter ‘J’ is denoted
by Js or J p depending on the location on the pile, i.e. damping offered by soil at the side
[skin] or toe. It should be noted that in static conditions Rt will be equal to Rs only. The
second part of Eq. (2) [Rs .J.v] is the instantaneous dynamic resistance offered by the soil. It
is obvious that the direction of this force will depend on soil compression or extension and
hence, the sign of this force should follow the sign of the velocity.
Several researchers through their limited laboratory tests opined that these damping
parameters generally do not vary linearly with the pile velocity. Researchers like Gibson
& Coyle (1968) and Heerema (1979) proposed an exponential law applied to the velocity
to calculate this dynamic resistance expressed in the form of, Rd = Rs .J.vn , for which ‘n’
depends on soil type and may be assumed as 0.18 for clay and 0.2 for sand. Goble and
Rausche (1976) proposed a damping factor Jc [= J.Ru /Z] which is much similar to Smith’s
viscous damping but it takes the effect of pile impedance [Z] in addition to soil ultimate
resistance [Ru ]. GRLWEAP recommends the use of Smith’s viscous damping approach in
case of tapered pile or while performing residual stress analysis.
A review of the soil damping parameters [Js & J p ] to be used for different subsoil condi-
tions proposed by various researches/consultants is presented in Table 1.

4. CASE STUDIES IN GULF OF THAILAND CLAY

Three case studies have been adopted from McDermott [MDR] project sites to analyze the
effect of remoulding and skin damping parameters on the driveability study of offshore
steel-tube piles driven through the clay dominated subsoil strata typically observed in Gulf
of Thailand. For the first case [Case Study 1 Gas Field GoT] the selected pile outer diameter
is 1,524 mm/2,350 mm penetrating up to 135 m with varying wall thickness [make-up] to
suit the hammer-pile system from the point of view of stress optimization and equipment
availability. A standard Menck MHU800S hammer was used to drive the pile with variable
efficiency increasing with depth. The soil strata generally consist of predominantly clay
layers with sands at deeper zones. The soil strata up to about 117.5 m consist of clay layers
with intermittent silt. Beyond this depth it was sand/silty sand until 131.3 m, interspersed
with thin layers of silt/clay, beyond which very stiff to hard clay layers continued to the
end of investigation depth of 150 m.
The second case study [Case Study 2 Oil Field GoT, Near Vietnam] is presented for com-
parison where the soil data is not predominantly clay. It was mixed with intermittent silt
and sand layers and the driven pile outer diameter was 2,133 mm driven to a target pene-
tration 126 m with the same Menck MHU800S hammer. The soil strata generally consist of
alternate layers of clay and sands. The soil strata up to about 55.8 m predominantly con-
sists of clay layers and beyond this depth is predominantly sand, interspersed with thin
layers of silt/clay, with the exception between 84 m and 95 m where the soil is stiff to hard
clay.
July 24, 2012 16:23: RPS: RPS-MOSS 2012

280 2nd Marine Operations Specialty Symposium

Table 1. Review of Soil Damping Factors Proposed.


Skin Damping, Js Toe Damping, Jp
sec/m sec/m Comments

Smith (1960) 0.16 0.49 All Soils


0.01–0.16 0.33–0.43 Lab tests – Granular Soil
Lowery et al (1969) 0.11 0.33 Best Correlation – Sand
0.33 0.98 Best Correlation – Clay
Forehand & Reese (1964) 0.16 0.16–0.65 Sand, Gravel
0.00 1.64–3.28
Poulos & Davis (1980) 0.23 0.66 Sand & Clay
Coyle et al 0.16 0.49 Sand
0.66 0.03 Clay
Lowery 0.33 0.49 Slit
0.16 0.33–0.65 Sand
GRLWEAP Manual 0.65 0.033–3.28 Clay
0.33–0.49 0.33–1.64 Slit
McCleland Memo 0.49 0.49 Sand
0.16–0.65 0.16 Js = f(Su) for Clay
Bowles (1977) 0.1–0.23 0.33–0.65 Sand
0.43–1.08 1.31–3.28 Clay
De Ruiter (1979) 0.16 0.49 Sand & Slit
0.66 0.03 Clay
Heerema (1980) 0.16 0.49 Sand
0.66 Clay
Beake & Sulcliffe 0.66 1.15 Carbonate Rock
Stevens (1982) 0.26 0.49 Using own method of SRD
0.10 0.49 assessment (Sand & Clay)
D’Appolonia 0.16 0.33 Gravel
0.16 0.49 Sand
0.33 0.49 Silty Sand & Clay
0.33 0.49 Calcareous Sand
Northern Field, Malaysia 0.16 0.49 Sand
0.30 0.49 Clay
Chermingat, Malaysia 0.20 0.50 Clay
Briaud & Tucker (1984) 0.16 0.49 Granular
Dolwin et al 0.52 0.20 Experiment & theory
Sulton et al 0.66 0.03 Clay
Tagaya & Heerema (1979) 0.49–0.65 0.49 Calcareous Clay
Van Luipen (1979) 0.66 0.36 Clay, some Sand
Semple et al (1981) 0.66 0.16 Clay
0.1–0.36 0.49–0.66 Clay
Roussel (1979) 0.23 0.49 Slit
0.26 0.49 Sand
Note: This is a collection of data available from different sources. The authenticity of all data presented
here is not completely verified by the author.

The third case study looks at in one of the Gas field in GoT, comprising a new design
with conditions similar to Case Study 1 location. The soil strata generally consist of pre-
dominantly clay layers and with sands at deeper zone. The soil strata up to about 4.3 m
consist of very soft clay layer followed by relatively firm to very stiff clay up to 26.7 m. This
July 24, 2012 16:23: RPS: RPS-MOSS 2012

Installation of Offshore Driven Piles – Regional Experience 281

Site for Case


Study 1 Site for Case
Site for Study 2
Case
Study 3

Figure 4. Case Study Investigation Locations.


is followed by alternate layers of medium to dense sand and very stiff clay inter-bedded
with silt until 64.8 m. Beyond this depth hard to very hard clay layers continued to the end
of the investigation depth of 149.7 m.
Figure 4 shows the location of the study fields in the Gulf of Thailand. The soil strata have
been schematically depicted in the corresponding driveability study curves in Figures 6
through 13.

4.1. Soil Resistance to Driving

The soil resistance to driving (SRD) is calculated using a similar procedure to that of
the axial capacity calculations following API 21st Edition method. The driving resistance
is determined either as a summation of external skin friction and overall end bearing
(plugged condition) or as a summation of external and internal skin friction and bearing
on the pile annulus (unplugged condition). The SRD is generally chosen as the unplugged
condition resistances with a check for the plugged condition as ultimate set-up. In the
present scenario, it has been assumed that with the usage of a thick and slanted driving
shoe, the effect of internal skin friction will be minimal and chances of plug formation will
be much less even in the case of a full soil set-up condition. However, both the effect of
external skin friction along with annular and fully plugged end bearing cases has been
considered in our present driveability study.
July 24, 2012 16:23: RPS: RPS-MOSS 2012

282 2nd Marine Operations Specialty Symposium

When a pile is driven in clay, there is a temporary reduction in the shear strength partly
due to remoulding and development of an undrained condition. This remoulded and
undrained condition is usually catered for by using set-up factors on the estimated SRD
in the clay layers. For sand and silt layers, there is no reduction and the capacities are
the same as those for the static conditions. The remoulded shear strengths are generally
assumed in the ranges from 20/30% to 60% of the design shear strength for the clays in
the Gulf area based on available experience. For the present analyses, separate SRDs have
been used to account for 30% [variation 20-40%] and 60% [variation 60-80%] external skin
frictional resistance for Clay in GRLWEAP with a set-up factor of 1.
Stoppages in pile driving generally result in the regaining of strength in clay layers and
hence the restart resistance is higher than the SRD for continuous driving. The gain in
strength is generally time dependent and hence longer delays generally result in a higher
restart SRD. In the absence of the site specific sensitivity data, the following values are used
for the prediction of the restart SRDs:

(a) Short delays (up to 8 hours) – 50% gain in strength [SRD shifts from 30% to 60% clay
Skin friction]
(b) Long delays (up to 24 hours) – 100% gain in strength [SRD shifts to 100% clay Skin
friction]

For delays beyond 24 hours, the shear strength will be assumed to be equal to static shear
strength.
Three sets of the following SRD have been used in the pile driveability analyses in the
Case Studies to ensure all driving conditions are covered:

(a) Continuous driving: 30% [may be 20–40%] skin friction resistance in Clay (external) –
API Unplugged case.
(b) Re-start after short delay: 60% [may be 60–80%] skin friction resistance in Clay (exter-
nal) – API Unplugged case.
(c) Re-start after long delay set-up: 100% skin friction resistance in Clay – Full Set-up Con-
dition, API Fully Plugged case.

In Case Study 2, Continuous driving scenario, SRD with 20% skin friction resistance in
clay has been used. In Case Study 3, a range of skin friction variation taken into account in
preliminary design of pile make-up and driveabilty.
Remoulding due to driving disturbance and time dependent strength recovery may be
best represented by a typical time rate of recovery curve as shown in Figure 5. Normally,
these curves are developed based on regional experience. An empirical procedure based
on experiments in Gulf of Mexico clays that include the effect of pile diameter and wall
thickness on the rate of consolidation and setup has been discussed in Bogard & Matlock
[1990].

4.2. Soil Dynamic Parameters for Wave Equation Analysis

The soil resistances as discussed above are modeled in GRLWEAP wave equation
analysis with both dynamic and static resistance wave equation parameters. The static
resistance component is displacement dependent and is defined by an ultimate strength
July 24, 2012 16:23: RPS: RPS-MOSS 2012

Installation of Offshore Driven Piles – Regional Experience 283

Figure 5. Rate of Increase in Axial Pile Capacity [Typical; Courtesy Somehsa Geosciences Pte.
Ltd.]

[Ru in Figure 3] and a quake [Q] or temporary soil compression usually related to soil
stiffness. The dynamic component is velocity dependent and known as viscous-damping
parameters which apparently increases the instantaneous soil resistance [indicated in
Figure 3]. The soil damping factors and quakes are taken based on typical values recom-
mended by researchers. Both the skin and toe quake values are considered as 2.5 mm and
the toe damping factors taken as 0.5 s/m for both sand and clayey subsoil. Since the site is
dominated by clayey subsoil, the most sensitive parameters would be skin damping fac-
tors. Roussel [1979] suggested that the skin damping in clay is a function of its strength
parameters [consistency] and he proposed a parameter range based on his experiments
[Table 1]. To study the sensitivity of this parameter, three cases have been considered in the
analysis as follows in Case Study 1:
Case 1 : Clay skin damping parameter taken as constant 0.65 s/m
Case 2 : Clay skin damping parameter taken as constant 0.30 s/m
Case 3 : Clay skin damping parameter taken as suggested by Roussel [1979]
Figure 6 shows the variation of skin damping plotted against the depth for these three
cases.

4.3. Results and Discussion

A set of driveability analyses have been performed for the study combining three sets
of SRDs and different skin damping parameters. The blow-count results are plotted in
Figure 7 [Case Study 1]. Generally, for wave equation analysis, where the subsoil condi-
tions are already well-defined, the dynamic soil parameters, i.e. soil damping coefficient
and quake should be taken based on the best-fit values recommended by earlier experi-
ence. The wave-equation analyses are quite sensitive to these parameters, particularly the
July 24, 2012 16:23: RPS: RPS-MOSS 2012

284 2nd Marine Operations Specialty Symposium

skin-damping parameters in clay. This is corroborated by the plots in Figure 7 which show
wide variation in all three cases of skin damping for the same SRDs. Actual in-situ values
of the parameters may vary widely depending on the type of subsoil in different stratum
and consistency.
Figure 7 also indicates the penetration at which the different add-ons are spliced. Because
of the welding delay, the restrike blow-counts are likely to increase at this penetration as
the SRD would increase from 30% to 60% skin-friction case or beyond up to full set-up.
Hence, the welding delay should be minimized as much as possible to avoid prolonged
hard driving or refusal in the worst case scenario. The hammer efficiency controls at dif-
ferent depths indicated in Figure 7 are principally to avoid an excessive run-down at the
initial stage and overstressing of the pile material.
Figures 8 and 9 incorporate the actual field driving record of the same piles plotted along
with the predicted or analyzed driveability blow-count results for comparison. The blow-
count and resulting soil resistances are monitored at site. The results clearly indicate the
sensitivity of skin damping in a clay dominated site. If we consider that the remoulded
skin frictions are predicted reasonably well to simulate different situations, e.g. continuous
driving or delayed driving as discussed earlier, the sensitivity of the results are evident
from Figure 9 where plots show a variable skin damping scenario. A skin damping in clay
closer to 0.3 sec/m or those suggested by Roussel [1979] are more realistic.
Figures 10 and 11 shown here are from the Case Study 2 in the same region when the
subsoil layers are not predominantly clay but more interspersed with intermittent silt and
sand layers. Because of the presence of the sand layers in between and towards the deeper
strata, the effects of clay skin damping parameters may not be so prominent in driveability
prediction. Nevertheless, from the actual field data it could be observed that with the sim-
ilar efficiency control the actual blow-count is closer to that predicted using a skin damp-
ing similar to that suggested by Roussel [1979] particularly at deeper strata [Figure 10].
The results also showed that the assumed SRD variation of 20–30% skin friction of clay
in the case of continuous driving and 60% for delayed driving holds good for the region
[Efficiency adjusted analyses; Figure 11]. A better comparison of SRD could be evaluated
by collecting instrumented pile monitoring data that can be obtained during pile driving
and subsequent back-analyses.
The following points are observed and to be noted from Figure 11:

• In Zone 1 indicated in the figure, the Clay has got more driving resistance than pre-
dicted /anticipated. During driving only 50% of the hammer efficiency was used in
this zone.
• In Zone 2 indicated in the figure, decrement of soil resistance due to driving after
restrike in the splicing zones are much slower and gradual than anticipated. This may
be due to gradual increment of hammer efficiency.

A third Case Study [Case Study 3 – Gas Field in GoT] taken from a new project in the
same region, shows preliminary analyses based on the available soil data. No field data
was available for this project. Figure 12 and Figure 13 show the predicted blow count for
several cases of efficiency variation and controlled efficiency driving case respectively.
Based on previous experience as well as per the soil consultants’ recommendations, only
Roussel [1979] damping parameters are used in the driveability analyses and predictions
July 24, 2012 16:23: RPS: RPS-MOSS 2012

Installation of Offshore Driven Piles – Regional Experience 285

Js [sec/m]
0 0.25 0.5 0.75
Driveability Analyses
0 With Variable Hammer Efficiency
Hammer: MHU800S

blows/m
0 200 400 600 800 1000
20 0
SRD 30%; Skin Damping Case 1
SRD 60%; Skin Damping Case 1
SRD FULL SET-UP; Skin Damping Case 1
SRD 30%; Skin Damping Case 2
16 SRD 60%; Skin Damping Case 2

Eff. 40%
SRD FULL SET-UP; Skin Damping Case 2
SRD 30%; Skin Damping Case 3
40 SRD 60%; Skin Damping Case 3
SRD FULL SET-UP; Skin Damping Case 3
32
P3 Add-on

Refusal Limit
60 48
Deoth [m]

Penetration [m]
64

80
P4 Add-on

Eff. 80%
80

100 96 P5 Add-on

112
120

Eff. 90%
128
Target Penetration 135m
Target Penetration 135m
140
144
Case 1, Js Clay 0.65s/m
Case 2, Js Clay 0.30s/m
Case 3, Js Roussel [1979]
160 Figure 7. Effect of Skin Damping on Driveabil-
ity Blow-count.
Figure 6. Variation of Skin Damping.

Driveability Analyses
Field Driving Records - All Data Compared Driveability Analyses
Hammer: MHU800S Damping Parameters as per Roussel [1979]
Hammer: MHU800S
blows/m
blows/m
0 200 400 600 800 1000 0 100 200 300 400 500 600 700
0
0

16 16
Efficiency = 12%
Low Efficiency Driving < 50%
Efficiency = 20%
P3 Add-on P2 Add-on
32 32

Efficiency = 25%

48 Efficiency = 40%
48

Efficiency = 20%
Efficiency = 60%
64 64
P3 Add-on
Penetration [m]

Efficiency = 70%
Penetration [m]

High Efficiency Driving > 90%


80 80
P4 Add-on
Efficiency = 80% P4 Add-on
Efficiency = 90%
96 96

112 112

Target Penetration
128
128
SRD 20% Eff. 70% SRD 20% Eff. 80%
Target Penetration
Efficiency = 95% SRD 20% Eff. 90% SRD 60% Eff. 70%
SRD 30% Eff. 70% SRD 30% Eff. 80% 144
144 SRD 30% Eff. 90% SRD 60% Eff. 70% SRD 60% Eff. 80% SRD 60% Eff. 90%
SRD 60% Eff. 80% SRD 60% Eff. 90%
FULL SET-UP Eff. 90% Pile E1 Field Data FULL SET-UP Eff. 70% Pile A1 Field Data
Pile B1 Field Data Pile E3 Field Data
Pile B3 Field Data 160
160

Figure 9. Comparison of Field Data Variable


Figure 8. Comparison of Field Data Single Effi-
Efficiency & Damping.
ciency & Damping.
Case Study 1 – Gas Field in GoT
July 24, 2012 16:23: RPS: RPS-MOSS 2012

286 2nd Marine Operations Specialty Symposium

Driveability Analyses Driveability Analyses


Damping Parameters as per Roussel [1979] Clay Skin Friction of 20% [Efficiency Adjusted as per Field]
Hammer: MHU800S
Hammer: MHU800S
blows/m
blows/m
0 100 200 300 400 500 600 700
0 100 200 300 400 500 600 700
0.0
0

16.0
16

Low Efficiency Driving < 50% P2 Added


P2 Add-on 32.0
32
Zone 1

48.0
48

64.0
64 P3 Added
P3 Add-on

Penetration [m]
Penetration [m]

Zone 2
High Efficiency Driving > 90% 80.0
80
P4 Added
P4 Add-on

96.0 Zone 2
96

112.0
112

Target Penetration
128.0
128
Pile A1 Field Data
SRD 20% Eff. 70% SRD 20% Eff. 80%
Clay Skin Damping as per Roussel [1979]
SRD 20% Eff. 90% SRD 60% Eff. 70%
144.0 Clay Skin Damping = 0.20 s/m
144
SRD 60% Eff. 80% SRD 60% Eff. 90% Clay Skin Damping = 0.30 s/m

FULL SET-UP Eff. 70% Pile A1 Field Data Clay Skin Damping = 0.65 s/m

160.0
160

Figure 10. Comparison of Predicted Observed Figure 11. Comparison of Observed & Pre-
Blow-count Blow-count. dicted with Skin Damping Variation.
Case Study 2 – Oil Field near Vietnam in GoT

Figure 13. Predicted Blow-count - Efficiency


Figure 12. Predicted Blow-count – All Cases Controlled
Case Study 3 – Gas Field in GoT
July 24, 2012 16:23: RPS: RPS-MOSS 2012

Installation of Offshore Driven Piles – Regional Experience 287

with different remoulded skin friction in clay for a continuous and delayed driving sce-
nario. The predicted SRDs have been used to account for a variation of 20–40% and 60–80%
of external skin frictional resistance for continuous and delayed driving scenarios respec-
tively. This should cover the predicted soil resistance during driving and the blow-counts
with a certain level of confidence and its fitness for the purpose.

5. CONCLUSIONS

On the basis of previous discussions and the case studies, the following conclusions can
be drawn with regards to driveability studies conducted on offshore driven piles in the
regional [Gulf of Thailand] subsoil.

• The optimum hammer selection is necessary for efficient, economic as well as safe driv-
ing. It is advisable to use a larger hammer with a low drop height but care should be
taken to avoid damaging the piles.
• Accuracy of driveability prediction is largely dependent on the accurate simulation/
representation of the in-situ soil parameters used to analyze the pile-soil interaction
problem. In addition to static soil parameters, the driveability study requires dynamic
parameters of the subsoil to which the analysis is quite sensitive.
• There are several methods proposed by different researchers available to calculate SRD.
Standard API RP 2A or ISO19902 methods with degraded clay skin friction variations
between 20–60% [of maximum static friction] to calculate lower and upper bound SRDs
[with a maximum upper bound of 80% in some conservative cases] are found to pro-
duce a reasonable prediction for subsoil strata in the Gulf of Thailand region. Field
observations and measurements/monitoring taken in some cases also corroborate this
hypothesis. Skin friction in sand is not normally degraded while determining SRD.
• Several researchers proposed various viscous-damping parameters based on their
experiments on different soil types. Limited dynamic laboratory tests indicated that
damping forces do not always vary linearly with pile velocity as is normally assumed
by the standard wave equation approach. In a standard laboratory or field testing sce-
nario, the hammer impact-velocity is typically in the range of 3 m/s whereas actual
pile driving is normally carried out with a ram velocity of around 5–6 m/s. This type
of variation would definitely affect the accurate interpretation of the damping model
that works in a non-linear domain.
July 24, 2012 16:23: RPS: RPS-MOSS 2012

288 2nd Marine Operations Specialty Symposium

• For sites dominated by clayey subsoil, the driveability prediction, subsequent accep-
tance criteria and selection of pile material and hammers are quite sensitive to skin-
damping parameters in clay. The careful selection of these parameters should be based
on regional experience. For a site with well-defined soil conditions, a constant clay
skin-damping parameter value between 0.2–0.3 sec/m is found to be suitable for the
Gulf of Thailand region. The dynamic parameters based on the clay consistency such as
proposed by Roussel [1979] may also be used with caution. For sand dominated sites,
as the skin-damping parameter for granular soil is generally lower, the driveability
analyses are not so sensitive to damping parameters.

ACKNOWLEDGEMENT

The Authors would like to thank McDermott Asia Pacific Engineering Management for
their support and encouragement in performing this study. Also the authors like to thank
the relevant Client personnel for their continuous help and review during the project phase
and successful completion. The team efforts for the members within McDermott are highly
appreciated. The authors are grateful to the reviewers for their constructive comments.

REFERENCES

1. Chen, R.P. and Chen, Y.M., 2000, Theoretical Study on Effect of Pile Shaft Resistance on Rebound
During Pile Driving, Proceedings of the 6th International Conference on the Application of Stress-wave
Theory to Piles, Brazil, pp 29–34.
2. Chen, R.P., Wang, S.F. and Chen, Y.M., 2003, Study on Pile Driveability with One dimensional
Wave Propagation Theory, Journal of Zhejiang University Science, V.4, No.6, pp 683–693, Nov.-Dec.,
2003.
3. Clough and Penzien, 1975, Dynamics of Structures, McGraw-Hill, Inc., New York.
4. Smith, E.A.L., 1960, Pile Driving Analysis by the Wave Equation, Journal of the Soil Mechanics and
Foundations Division, Transaction of ASCE, Paper No. 3306, Vol. 127, 1962, Part I, pp 1145–1193.
5. Rausche, F., Goble, G.G. and Likins, G., 1992, Investigation of Dynamic Soil Resistance on Piles
Using GRLWEAP, Application of Stress-Wave Theory to Piles, F.B.J. Barends (ed.), Balkema, Rotter-
dam, ISBN 90 5410 0826, pp 137–142.
6. API RP 2A, 21st Edition, 2000, Recommended Practice for Planning, Designing and Constructing Fixed
Offshore Platforms – Working Stress Design, American Petroleum Institute.
7. ISO19902:2007, Petroleum and Natural Gas Industries – Fixed Steel Offshore Structures.
8. GRLWEAP, Wave Equation Analysis of Pile Driving, GRL Engineers, Inc., 2005.
9. Alm, T. and Hamre, L., 1998, Soil Model for Driveability Predictions, Offshore Technology Confer-
ence, Houston, Paper No. OTC 8835, May, 1998.
10. Toolan, F.E. and Fox, D.A., 1977, Geotechnical Planning of Piled Foundations for Offshore Plat-
forms, Proceedings of Institution of Civil Engineers Part I, May 1977, Paper No. 7996.
11. Stevens, R.M. and Wiltsie, A.W. and Turton, H.T., 1982, Evaluating Pile Driveability for Hard
Clays, Very Dense Sand and Rock, Offshore Technical Conference, 1982, OTC Paper No. 4205.
12. Semple, R.M. and Gemeinhardt, J.P., 1981, Stress History Approach to Analysis of Soil Resistance
to Pile Driving, OTC, Houston, Vol. 1 pp 165–172, 1981, Balkema.
13. Gibson, G.C. and Coyle, H.M., 1968, Soil Damping Constants Related to Common Soil Proper-
ties in Sands and Clays, Research Report No. 125-1, Texas Transportation Institute, Texas A & M
University.
14. Heerema, E.P., 1979, Relationships between Wall Friction, Displacement Velocity, and Horizontal
Stress in Clay and in Sand, for Pile Driveability Analysis, Ground Engineering, Vol. 12, No. 1.
15. Goble, G.G. and Rausche, F., 1976, Wave Equation Analysis of Pile Driving – Program Manuals,
Department of Transportation, Report No. FHWA IP-76-14-3.
July 24, 2012 16:23: RPS: RPS-MOSS 2012

Installation of Offshore Driven Piles – Regional Experience 289

16. Uto, K., Fuyuki, M. and Omar, H., 1992, New Development of Pile Driving Management
System. Proceedings of 4th International Conference on the Application of Stress Wave Theory to Piles,
The Hague, pp 351–356.
17. Weele, A.F. and Schellingerhout, A.F.G., 1994, Efficient Driving of Precast Concrete Piles, Proceed-
ings of Conference Development in Geotechnical Engineering, Bangkok, pp 1–8.
18. Chen, Y.M., Chen R.P., Wu, S.M. and Van Weele, A.F., 1997, Determining Static Resistance of
Pile Toe during Driving with Acceleration of Pile Top, Chinese Journal of Geotechnical Engineering,
19(6):16–21.
19. Roussel, H.J., 1979, Pile Driving Analysis of Large-Diameter, High-Capacity Offshore Pipe Piles,
PhD Thesis, Department of Civil Engineering, Tulane University, New Orleans, Louisiana.
20. Lowery, L.L., T. J. Hirsch, T. C. Edwards, H. M. Coyle and C. H. Samson, Jr., 1969, Pile Driving
Analysis State of the Art, TTI Research Report No. 3313 (Final).
21. Forehand, P.W. & Reese, J.L., 1964, Prediction of pile capacity by the wave equation, Journal of
Soil Mechanics and Foundations. ASCE, March 1964, No. SM2, pp 1–25.
22. Coyle, H.M. & Gibson, G.C., 1970. Empirical damping constants for sands and clays, Journal of
Soil Mechanics and Foundation, ASCE, Vol 96, No. SM3, pp 949–965.
23. Poulos, H. G. and Davis, E. H., 1980, Pile Foundation Analysis and Design, John Wiley and Sons.
24. Briaud, J.L. and Tucker, L.M., Coefficient of Variation for In Situ Tests in Sand, ASCE Symposium
on Probabilistic Characterization of Soil Properties, Atlanta, April 1984.
25. Bowles, J.E., Foundation Analysis and Design, McGraw Hill, 1977
26. de Ruiter, J. and Beringer, F., 1979, Pile Foundations for Large North Sea Structures, Marine
Geotechnology, 3, No. 3, pp. 267–314.
27. Heerema, E.P., 1980, Predicting pile driveability: Heather as an illustration of the ”friction
fatigue” theory, Ground Engineering, 13, 15–37.
28. Tagaya, K., Heerema, E. P., Uchino, T. and Kusaka, T., 1979, Pile driveability test on actual
offshore platform in calcareous clay for Qatar NGL offshore project, OTC 3440. Proceedings, 11th
Annual Offshore Technology Conference. Vol. 2, p. 713–16.
29. Van Luipen, P. and Jonker, G., 1979, Post Analysis of Full-scale Pile Driving Test, Proceedings of
International Conference on Numerical Methods in Offshore Piling, ICE, London, pp 43–52.
30. Health & Safety Executive (2001), “A study of pile fatigue during driving and in-service and of
pile tip integrity”, HSE Books, Sudbury, Offshore Technology Report, 2001/018.
31. Schneider & Harmon, 2010, Analyzing Driveability of Open Ended Piles in Very Dense Sands,
DFI Journal, Vol. 4, No. 1 August 2010.
32. Lehane, B.M., Jardine, R.J., Bond, A.J. and Frank, R. (1993), Mechanisms of Shaft Friction in Sand
From Instrumented Pile Tests, Journal of Geotechnical Engineering, 119(1), 19–35.
33. Lehane, B.M., Schneider, J.A. and Xu, X., (2005), The UWA-05 Method for Prediction of Axial
Capacity of Driven Piles in Sand, Proceedings of the International Symposium on Frontiers in Offshore
Geomechanics, Perth, Australia, 683–690.
34. Lehane, B.M., Schneider, J.A. and Xu, X., (2007), Development of UWA-05 Design Method for
Open and Closed Ended Driven Piles in Siliceous Sand, Contemporary Issues in Deep Foundations,
ASCE GSP 158, 1–10.
35. Randolph, M.F., 2003, Science and Empiricism in Pile Foundation Design, Geotechnique, 53(10),
847–875.
36. Randolph, Mark, Cassidy, M., Gourvenec, S. and Erbrich, C. (2005), “Challenges of Offshore
Geotechnical Engineering”, Osaka 2005.
37. Bogard, J.D. and Matlock, H., 1990, Application of Model Pile Tests to Axial Pile Design,
Proceedings, 22nd Annual Offshore Technology Conference, Houston, Texas, Vol. 3, pp. 271–278.
38. Schneider, A. and Harmon, Ivy A., 2010, Analysing Driveability of Open Ended Piles in Very
Dense Sands, DFI Journal, August, 2010, pp. 32–44.

View publication stats

You might also like