You are on page 1of 12

Static Response of Monopile to Lateral Load

in Overconsolidated Dense Sand


Weichao Li 1; Bitang Zhu 2; and Min Yang 3

Abstract: Large diameter, rigid monopile foundations have been extensively used in the fast-growing offshore wind energy industry over the
last two decades. In view of the offshore environment, lateral response of the monopile usually governs its design. Even though several
Downloaded from ascelibrary.org by University of Newcastle on 03/12/17. Copyright ASCE. For personal use only; all rights reserved.

approaches have been recommended based on small-scale laboratory tests, there is no widely accepted method for the design of monopiles
under lateral loading. Conversely, field testing on large-diameter prototype monopiles is normally impractical due to the high demand on the
capacity of loading equipment. For these reasons a series of field lateral loading tests on reduced-scale monopiles were conducted at a dense
sand test bed site. The model monopiles had similar aspect ratios of pile length to diameter to those used in the offshore wind farm projects,
but were smaller in scale. Experimental p-y curves were derived using the measured monopile responses of the lateral load tests, and a
distinctive failure model was presented for monopiles in the overconsolidated dense sand deposit. Comparison of lateral responses of mo-
nopile between the measured and predicted by two current p-y models showed the shear force at the pile tip plays an important role
and should be accounted for in the design of laterally loaded rigid monopiles. Finally, a refined p-y model for laterally loaded rigid monopiles
in overconsolidated dense sand was recommended and calibrated. DOI: 10.1061/(ASCE)GT.1943-5606.0001698. © 2017 American Society
of Civil Engineers.
Author keywords: Monopile; Field tests; Overconsolidated dense sand; p-y curves; Pile tip shear force.

Introduction tangent hyperbolic (e.g., Murchinson and O’Neill 1984), hyper-


bolic curves (e.g., Georgiadis et al 1992), and power law functions
As one of the most efficient foundation to resist lateral loading, pile (e.g., Wesselink et al 1988). Among them, the tangent hyperbolic
foundations have been widely used in the offshore gas and oil in- p-y models are recommended by several design codes for laterally
dustry and investigated extensively since the 1960s. Multiple loaded piles (e.g., API 2005; DNV 2007). For a given site, the
methods were proposed to aid the laterally loaded pile design in construction of hyperbolic p-y models entails the determination
the past decades. One of the most commonly used methods is of two key parameters, i.e., the initial stiffness (K i ) of p-y curves,
the subgrade reaction method (Hetenyi 1946), which simplifies the and ultimate soil resistance (pu ¼ A × put , where A is the empiri-
pile as an elastic beam in the ground and the ground soil is repre- cal depth factor and put is the theoretical ultimate soil resistance).
sented by a series of uncoupled springs along the pile’s embedded The initial stiffness is usually assumed to increase linearly
length. A number of elastic methods have been proposed to deter- with depth and depend on ground soil relative density, but the
mine the stiffness of these uncoupled springs (e.g., Yoshida and increasing rate (ni , termed constant of initial stiffness) varies be-
Yoshinaka 1972; Shadlou and Bhattacharya 2014). To account for tween different studies (API 2005; Georgiadis et al 1992). Addi-
the nonlinear effect of ground soil a tentative model with nonlinear tionally, different studies recommended different equations to
degradation of soil stiffness with increasing strain/displacement calculate the ultimate soil resistance (Reese et al. 1974; Klinkvort
was introduced by McClelland and Focht (1956) and then devel- and Hededal 2014). Additionally, cone penetration test (CPT)
oped for different soil types (e.g., Matlock 1970; Reese et al. 1974), based p-y models (Li et al. 2014) and laboratory stress-strain
which is usually called force-displacement model, i.e., the p-y curves based p-y models (Bouzid et al. 2013) have also been
method, where p is soil lateral reaction (in force per unit length) developed.
and y is soil lateral displacement or the pile deflection at the point With the fast development of the offshore wind energy industry,
under interest. monopiles for which the loading conditions could be more clearly
To date, for laterally loaded pile in sand, numerous p-y models defined than pile groups or other foundations have been increas-
were proposed, and most of them are expressed with functions of ingly used in real-world applications. The monopiles usually have
a smaller slenderness ratio (RLD) of embedded length (Lem ) to the
1
Assistant Professor, Dept. of Geotechnical Engineering, Tongji Univ., outer diameter (D), i.e., RLD ¼ Lem =D, and are classified as short-
Shanghai 200092, P. R. China (corresponding author). ORCID: http://orcid. rigid piles (Leblanc et al. 2010). Due to the small value of RLD, the
org/0000-0001-9992-9300. E-mail: WeichaoLI@tongji.edu.cn monopiles behave as rigid piles with one point of zero displacement
2
Principal Engineer, NOMA Consulting Pty Ltd., Level 1, 530 along its embedded length (Reese and Van Impe 2001), and rota-
Little Collins St., Melbourne, VIC 3000, Australia. E-mail: bitang. tion of pile shaft cannot be neglected. Unlike long-flexible piles
zhu@noma-consulting.com with attenuating displacement with depth, the monopiles may have
3
Professor, Dept. of Geotechnical Engineering, Tongji Univ., Shanghai
relatively large displacement at the pile tip. Therefore, lateral
200092, P. R. China. E-mail: yangmin@tongji.edu.cn
Note. This manuscript was submitted on September 22, 2015; approved shear force at the pile tip would be developed for the short-rigid
on December 15, 2016; published online on March 11, 2017. Discussion monopiles (e.g., Reese and Van Impe 2001; Abdel-Rahman and
period open until August 11, 2017; separate discussions must be submitted Achmus 2005).
for individual papers. This paper is part of the Journal of Geotechnical and For the static design of laterally loaded monopiles, p-y models
Geoenvironmental Engineering, © ASCE, ISSN 1090-0241. derived from field tests on the long-flexible piles are still used on a

© ASCE 04017026-1 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., -1--1


100
provisional basis. Even though some successful design cases with
these models have been reported (Klinkvort and Hededal 2014),
the applicability of existing p-y models to laterally loaded monop- 80
ile design has been questioned (Jardine et al. 2012). Based on

Passing (%)
comprehensive finite-element modelling, Abdel-Rahman and 60
Achmus (2005) and Lesny and Wiemann (2006) found that the
linear distribution of initial stiffness recommended by American 40
Petroleum Institute (API) guidelines led to an overestimation of z=1.75m
initial stiffness at greater depths. Klinkvort and Hededal (2014) z=2.85m
20
performed a series of centrifuge tests on model monopiles and z=4.50m
found that the constant for determination of ultimate soil resis-
tance is clearly site-dependent. Even though several studies re- 0
0.01 0.1 1
vealed that the lateral shear force at short-rigid pile (Poulos
Sieve size (mm)
Downloaded from ascelibrary.org by University of Newcastle on 03/12/17. Copyright ASCE. For personal use only; all rights reserved.

and Hull 1989) tip plays an important role, no shear force-


displacement models have been proposed. In addition, most
Fig. 1. Particle size distribution (adapted from Kirwan 2015)
current studies on laterally loaded monopiles are based on
small-scale laboratory tests or numerical modelling. Therefore,
a further verification of these studies with field tests on larger
or full-scale monopiles is urgently needed. 0
σmax
The objective of this paper is to present and analyze a series of OCR ¼ ð1Þ
σv0
field lateral load tests on reduced scale monopiles in overconsoli-
dated dense siliceous sand. The test results are carefully examined,
and some responses of monopiles are further discussed. Experi- K 0 ¼ ð1 − sin ϕcv Þ × OCRsin ϕcv ð2Þ
mental p-y curves and lateral shear force-displacement trend at pile
tip are derived from the analysis of the tests. Finally, a refined p-y where σmax0 and σv0 = effective preconsolidation stress and in situ
model is recommended for the design of laterally loaded monopiles vertical effective stress, respectively.
driven in dense sand. The test program presented in Table 1 was designed to inves-
tigate the response of short-rigid monopiles subjected to lateral
static loading in the larger-scale that is found in the field. Two sets
Test Description of static lateral loading tests (termed PS1 and PS2) were conducted
6 and 19 weeks after pile driving, respectively. In addition, another
The lateral loading tests on reduced-scale model monopiles were two sets of cyclic loading tests (termed PC1 and PC2) on piles with
conducted in Blessington Co. Wicklow, 25 km southwest of Dublin the same dimension were conducted at this site, which has been
city, Ireland. The site has been used extensively for pile testing re- reported by Li et al. (2015).
search and details of the ground conditions can be found in Gavin The tests were conducted on two 3-m long open-ended steel
and O’Kelly (2007), Tolooiyan and Gavin (2011), Gavin et al. tubes with an outer diameter of 340 mm and wall thickness of
(2014), Li et al. (2015), and others. In summary, the uniform sand 14 mm [Fig. 3(a)]. Both test piles were instrumented with series
deposit confirmed by extensive excavations at the site is heavily a of strain gauges. To prevent the strain gauges from being dam-
overconsolidated because of the glacial history and recent exca- aged during the pile installation especially in such a dense site,
vations of the area. The maximum preconsolidated pressure is two channels on each pile were welded on the diametrically
approximately 800 kPa, and are estimated from oedometer tests. opposite sides of each pile along the loading direction as shown
Sand replacement tests suggest the in-situ relative density ðDr Þ is in Fig. 3(b). Each 2.3-m long channel was composed of two 16 ×
close to 100%, and the unit weight of the soil is 20 kN=m3 and rel- 16-mm square steel bars and a 40-mm wide by 3.5-mm thick top
atively constant with depth. The water table is approximately 15 m plate. The model monopile for each set of tests was driven down to
below the ground surface, and above the ground water level the de- 2.2 m below the ground level with a 5-t Juntann hammer, giving
gree of saturation is approximately 70%. The particle size distribu- RLD ¼ 6.5 for both test piles.
tion is shown in Fig. 1 and the mean particle size (D50 ) is between To capture the moment profiles of the pile during the lateral
0.1 and 0.15 mm, and the fine content (percentage of clay and silt loading tests, 10 levels=pairs of 350 Ω strain gauges (serial number
particles) is between 2 and 14%. The average values of (D60 ) and 1-LE11-3/350Z) were bonded along the pile embedded depth at a
(D10 ) of the sand particles are 0.1−0.17 mm and 0.07 mm, respec- constant spacing of 0.23 m [Fig. 4(a)]. The lower pair of strain
tively. Triaxial compression tests on reconstituted Blessington sand gauges was located at 0.10-m above the pile tip. The pair of strain
samples with Dr close to 100% indicate the constant volume friction gauges at each level/cross section could measure the maximum
angle ðϕcv Þ is 37° and the peak friction angle ðϕÞ decreases from 54° tension and compression strains separately. Each strain gauge
at a 1-m depth to 42° at a depth of approximately 5 m. was wired in a three wire quarter bridge configuration.
Multiple CPTs were performed at the pile testing site. The Pile deflections were measured at three height levels by Linear
maximum, minimum, and averaged values of cone tip resistance Variable Differential Transformers (LVDTs) [Fig. 5(a)], which
(qc ) and sleeve friction (fs ) are plotted in Figs. 2(a and b), were fixed on the reference beam. For PS2 an additional six incli-
which show that qc values increased from 10 megapascals nometers [Fig. 5(b)] were mounted at the same levels of the LVDTs
(MPa) close to the ground level to approximately 20 MPa at to measure the pile rotations.
6 m below ground level. Additionally, according to laboratory A hydraulic jack was used to apply lateral load and mounted
tests results previously presented the profiles of overconsolidation at 0.4-m above ground level. A load cell was set up between
ratio (OCR) and lateral earth pressure coefficient at rest (K 0 ) the hydraulic jack and the test pile to monitor the applied lateral
estimated by Eqs. (1) and (2) are also presented in Figs. 2(c load. The test set-up of PS2 is shown in Fig. 6. The lateral
and d), respectively load is applied by stages and each load increment is 10 kN and

© ASCE 04017026-2 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., -1--1


Cone tip resistance, qc (MPa) Sleeve friction, fs (MPa)
0 10 20 30 0 0.1 0.2 0.3 0.4
0 0

1 1 fs,mean
qc,mean

2 qc,max 2
Depth, z (m)

Depth, z (m)
3 Fitted qc with 3
power function

fs,max
Downloaded from ascelibrary.org by University of Newcastle on 03/12/17. Copyright ASCE. For personal use only; all rights reserved.

4 4

5 qc,min 5
fs,min
6 6
(a) (b)

Over-consolidation ratio, OCR Lateral earth pressure coefficient at rest,


K0
0 50 100 150 0 5 10 15 20
0 0

1 1

2 2
Depth, z (m)
Depth, z (m)

3 3

4 4

5 5

6 6
(c) (d)

Fig. 2. Profiles of (a) cone tip resistance; (b) sleeve friction; (c) overconsolidation ratio; (d) earth pressure coefficient at rest

applied and each load increment held for 5 minutes. The detailed
Table 1. Test Schedule
loading schedule is summarized in Table 1.
Number The applied lateral load and induced bending strains, pile
of load- Maximum deflections, and rotations were recorded concurrently at a sample
Pile Date of Date of unloading load rate of 10 readings per second.
number installation loading cycles applied (kN)
Two main scale effects have been considered in the lateral
PS1 April 27, 2011 June 8, 2011 3 40 loading tests on scaled monopiles (Li et al. 2015). First, the test
June 21, 2011 4 40 monopiles’ geometric dimensions and applied loads need to be
July 8, 2011 2 110 compatible with those for a typical monopile foundation used in
PS2 May 28, 2012 October 11, 2012 4 110
the offshore wind turbine. Assuming a typical monopile with outer
diameter of 6 m, the test monopile with a diameter of 0.34 m
approximately represents a 1/18 model scale. For the magnitude
and height (e ¼ 0.4 m) of applied lateral loads in the test, the
approximately 10% of the estimated ultimate lateral load capacity equivalent moment at the ground level is in the range of
(approximately 70−120 kN), which was predicted with the API 163−280 MN · m, which is compatible with the loads applied
method and finite-element modelling (Doherty et al. 2012). At by offshore wind turbine (Doherty and Gavin 2012). Secondly,
some load stages several numbers of loading-unloading were the effect of reduced stress level on the strength and stiffness of

© ASCE 04017026-3 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., -1--1


Loading direction

Weld joint

340 mm

312 mm
Steel plate

Steel bar
Downloaded from ascelibrary.org by University of Newcastle on 03/12/17. Copyright ASCE. For personal use only; all rights reserved.

(a) (b)

Fig. 3. Cross section of model monopiles: (a) uninstrumented; (b) instrumented

(Unit: cm)
Pile Wall

Ground Level

Channel

Strain Gauge

(a) (b)

Fig. 4. Scaled monopile instrumented with strain gauges: (a) arrangement; (b) photograph

ground soil has been considered. The tip resistances of CPT testing Pile Head Response
in a range of 10−20 MPa are comparable to typical soil condition
The measured model monopile displacement and rotation at the
encountered by monopiles for offshore wind farms, and many re-
ground level of PS1 and PS2 are plotted in Fig. 8 (no rotation mea-
searchers (Lehane 1992; Gavin and O’Kelly 2007; Li et al. 2014)
surements were made for PS1), which shows that as the applied
have established the correlations between pile response and the tip
load increased, the response of both test piles transited from linear
resistance of CPT.
response to nonlinear and reached to a plateau which indicates the
ultimate capacity reached. The displacement needed to mobilize the
ultimate load is approximately 17 mm (0.05D) at ground level. As
Test Results and Analysis
the ground water table is approximately 15-m below the ground
The driving record of PS2 is shown in Fig. 7. It shows that the level and the sand is not fully saturated, the pore water pressure
number of blows per 250 mm driving [Fig. 7(a)] increases almost effect on the pile response during loading should be minimal.
linearly with the driving depth. This trend agrees well with that of
qc values measured from CPTs, see Fig. 2(a). Correspondingly, the
Geometric Dimension of Failure Wedge
driving depth of each blow almost linearly decreases in the semi-
logarithmic plot [Fig. 7(b)] after a sharp reduction within approx- At the initial loading stages for both PS1 and PS2, small hairline
imately 1 m below the ground level. cracks in the sand were visible that radiated outwards from the pile.

© ASCE 04017026-4 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., -1--1


Blows/250 mm Driving depth/Blow (mm)
0 5 10 15 20 1 10 100 1000
0 0

1 1

Depth, z (m)

Depth, z (m)
2 2
Downloaded from ascelibrary.org by University of Newcastle on 03/12/17. Copyright ASCE. For personal use only; all rights reserved.

3 3
(a) (a) (b)

Fig. 7. Test pile driving records of PS2: (a) blow counts per 250-mm
depth; (b) driving depth per blow

120

Lateral load, F0 (kN)


100

80

60

40 PS1
20 PS2

0
(b) 0 10 20 30
(a) Lateral displacement, y0 (mm)
Fig. 5. (a) LVDT; (b) inclinometer used in the pile tests
50
Moment, M0 (kN.m)

40

30

20
PS2
10

0
0.0 0.2 0.4 0.6 0.8 1.0
(b) Rotation, 0 (degree)

Fig. 8. Measured pile head response of (a) load-displacement; (b) mo-


ment-rotation

Fig. 6. Lateral loading test set-up for PS2


observed from Reese’s test. Interestingly Bowman (1958) had
found that, as referred by Reese et al. (1974), α is probably a func-
As the load increased, a gap approximately 15-mm wide and 1-m tion of the void ratio of the sand, with values ranging from ϕ=3 to
deep was developed behind the pile. This was coupled by sharply ϕ=2 for loose sand to ϕ for dense sand. This infers that the obser-
defined cracks radiating from the pile diameter [Fig. 9(a)]. The vation from the Blessington lateral loading tests in this particularly
same phenomenon was observed during the lateral cyclic loading overconsolidated dense sand deposit agrees well with Bowman’s
tests on PC1 and PC2 (Li et al. 2015). Fig. 9(b) shows the wedge conclusion, which was generally derived from model test results
failure model near the ground surface, which was suggested by Re- with a small flat plate in sand.
ese et al. (1974) for normally consolidated medium dense sand. The
geometry of this wedge is defined by the pile diameter, the depth of
Derivation of Pile-Soil Response
the wedge z, and the angle of α and β. Fig. 9(a) shows that α was
approximately 54°, which is the value of peak friction angle at a 1- The bending moment (M) profiles for PS2 are plotted in Fig. 10(a),
m depth and noticeably different from the value of ϕ=2 which shows that there is no negative bending moment along the

© ASCE 04017026-5 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., -1--1


Downloaded from ascelibrary.org by University of Newcastle on 03/12/17. Copyright ASCE. For personal use only; all rights reserved.

(a)

(b)

Fig. 9. (a) Cracks developed around PS1; (b) wedge failure model recommended by Reese et al. (1974)

pile depth. To derive the p-y curves from a fully instrumented test integration of curvature/bending moment profile could produce a
pile a conventional method is to use a mathematical function reliable displacement profile; however, direct differentiation of
(e.g., nth order polynomial) to fit the measured curvature (κ) or the curvature/bending moment profile leads to the amplification
bending moment profile. Based on the fitted curvature function, of measurement errors and may give unacceptable shear force and
the deflection profile can be backcalculated with the double integra- soil reaction profile. To overcome these uncertainties, the procedure
tion of curvature profile with respect to depth as is shown in Eq. (3): recommended by Nip and Ng (2005) is modified by assuming the
Z Z  profiles of bending moment and soil reaction are represented by
y¼ κ · dz · dz ð3Þ Eqs. (6) and (7). The corresponding soil reaction, pile rotation,
and deflection can be expressed as Eqs. (8)–(10)

Conversely, the first and second derivatives of the moment pro- a 4 b 3 c 2


M¼ z þ z þ z þ F0 · z þ M 0 ð6Þ
file with respect to depth are used to derive the shear force (Q) 12 6 2
profile and soil reaction (p) profile, respectively:
a b
dM Q ¼ z3 þ z2 þ cz þ F0 ð7Þ
Q¼ ð4Þ 3 2
dz
p ¼ az2 þ bz þ c ð8Þ
d2 M
p¼ ð5Þ 2 3
dz2 a 5 b 4
þ 24 z þ 6c z3 þ F20 z2 þ M0 z 7
6 60 z
θ¼4 5 þ C0 ð9Þ
Previous researchers (Matlock and Ripperger 1956; Nip and EI
Ng 2005; Yang and Liang 2007) have concluded that the double

© ASCE 04017026-6 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., -1--1


Bending moment profiles, M Shear force profiles, Q (kN)
(kN.m)
0 30 60 90 -90 -60 -30 0 30 60 90 120
0.0 0.0

0.5 0.5

Depth, z (m)

Depth, z (m)
1.0 1.0
Downloaded from ascelibrary.org by University of Newcastle on 03/12/17. Copyright ASCE. For personal use only; all rights reserved.

1.5 30kN,Fit 1.5


30kN,Measured
60kN,Fit 30kN
60kN,Measured
60kN
2.0 110kN,Fit 2.0
110kN,Measured
110kN

(a) (b)

Soil reaction profiles, p (kN/m) Deflection profiles, y (mm)


-200 -100 0 100 200 -10 0 10 20
0.0 0.0

0.5 30kN 0.5


60kN
110kN
Depth, z (m)

Depth, z (m)

1.0 1.0

1.5 1.5
30kN
60kN
110kN
2.0 2.0

(c) (d)

Fig. 10. Derived profiles of (a) bending moment; (b) shear force; (c) soil reaction; (d) displacement

2 3
a 6 b 5 c 4 Depth of Rotation Point
6 360 z þ 120 z þ 24 z þ F60 z3 þ M20 z2 7
y¼4 5 þ C0 z þ y0 Figs. 10(c and d) show that the values of rotation depth zr (Fig. 11)
EI
ranged from 1.60 (0.73Lem ) to 1.87 m (0.85Lem ) below the ground
 a 6 b 5 c 4
 level, with the smaller rotation depth for the highest load level
360 zr þ 120 zr þ 24 zr þ F60 z3r þ M20 z2r
þ y0 of 110 kN. The location of the rotation point agrees well with
EI
C0 ¼ ð10Þ the conclusion given by Klinkvort and Hededal (2014) that as
zr the applied loading increases the rotation depth decreased from
where a, b, and c = constants to be determined; F0 and M0 = the maximum value of approximately 0.87Lem to a constant value
applied lateral load and bending moment at ground level, respec- of 0.70Lem . In addition to the effect of load magnitude, cyclic load-
tively. According to the measured bending moment data, the least- ing also makes the rotation point move up (Peralta 2010; Li et al.
square method is employed to determine the constants in these 2015); for example, the depth of rotation point of PC2 (Li et al.
equations. In addition, compatibility between soil reaction and de- 2015) decreased by approximately 10% (from 0.79Lem to 0.71Lem )
flection profile (i.e., the depth of zero lateral soil reaction is equal to after approximately 2,500 load cycles. The mechanism behind the
that of zero lateral displacement of pile) is imposed. The derived rotation depths at a relative narrow range may be attributed to the
profiles of shear force, soil reaction, and deflection are plotted in mobilization of the soil resistances primarily at the upper and
Figs. 10(b–d). bottom parts of the pile.

© ASCE 04017026-7 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., -1--1


Shear Force at Pile Tip
A further study on the lateral displacement and shear force in the
Figs. 10(b and d) indicates that a shear force should not be ignored
at the pile tip (Fig. 11) and it increases with the increasing in pile tip
displacement. This agrees with the findings by other researchers
(Reese and Van Impe 2001) that shear force (Qb ) at the tip of
short-rigid pile could be developed due to the nonnegligible dis-
placement at the tip and expressed as a function of pile tip defection
(yb ). The derived shear force (Qb ¼ Qz¼Lem ) and corresponding
deflection at the pile tip are plotted in Fig. 12. It shows that as
the lateral displacement (yb ) increases, the shear force (Qb ) in-
creases to a maximum value of approximately 35 kN and then drop
to an almost constant value of 32 kN, which suggests the pile tip
Downloaded from ascelibrary.org by University of Newcastle on 03/12/17. Copyright ASCE. For personal use only; all rights reserved.

shear force and displacement exhibit softening behavior. As the


ground soil is overconsolidated dense sand and the depth of pile
tip is 2.2 m, the tip shear force reaches the peak value and then
drops to the residual value at higher tip movments.
To investigate the effect of shear force at the pile tip, p-y models
proposed by API (2005) and Li et al. (2014) are employed to cal-
culate the lateral responses of PS2, including pile deflection at
ground level, pile head rotation, and bending moment distribution.
For each p-y model two scenarios are considered: one without tip
shear force (Qb ¼ 0), and the other with tip shear force (Qb ≠ 0)
derived from the loading test results (Fig. 12). The predicted re-
sponses against the measured data are shown in Fig. 13. It shows that:
1. The pile responses with tip shear force are stiffer than those
without tip shear force;
2. By comparison with the measured bending moment distribution,
the p-y curves with tip shear force considered give a better pre-
diction than those with tip shear force ignored, especially close
to the pile tip. Therefore, for monopiles the shear force at the
pile tip should be taken into account, which agrees with the re-
commendation given by Abdel-Rahman and Achmus (2005).
The shear force will be considered in the following analysis; and
3. The p-y model proposed by Li et al. (2014) with tip shear force
considered offers better prediction than the API p-y model,
especially for the high loading levels. Therefore, caution should
be exercised when p-y model originated for flexible piles is
employed to the design of rigid monopiles.

Derived p-y Curves


Fig. 11. Sketch of a laterally loaded monopile
Based on the derived profiles of the soil lateral reaction (p) and
deflection (y) at each load level, the p-y curve at each depth along
the test pile’s embedded length can be plotted. Fig. 14 shows the
comparison of normalized p-y curves between measurements from
PS2 and the recommended p-y curves by API (2005), in which the
p is normalized by γD2 and the y is normalized by D and expressed
as a percentage. It shows that:
1. The API p-y curve is generally stiffer at initial stage and softer at 40
Shear force at pile tip, Qb (kN)

larger displacement, which agrees with the previous findings


(Yan and Byrne 1992). 30
2. The initial stiffness of measured p-y curves increase with depth,
but do not increase significantly. This insignificant change with
depth may result from this overconsolidated dense sand deposit 20
site (Fig. 2). By comparison, as the depth increases API gradu-
ally overestimates the initial stiffness. This is in line with the 10
current findings (Sorensen et al. 2010; Klinkvort and Hededal
2014) that the API method overestimates the initial soil stiffness
at a large depth. 0
0 1 2 3 4 5 6
3. Except for soil near the ground level, the measured soil reaction
values didn’t reach the plateau (i.e., the ultimate lateral soil re- Pile tip displacement, yb (mm)
action) in the tested monopiles. However, the API method
Fig. 12. Lateral shear force versus displacement at pile tip
appears to underestimate the soil’s ultimate lateral soil reaction,

© ASCE 04017026-8 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., -1--1


Downloaded from ascelibrary.org by University of Newcastle on 03/12/17. Copyright ASCE. For personal use only; all rights reserved.

(a)

(c)
M

(b)

Fig. 13. Pile tip shear force effect on pile response: (a) load-displacement at ground level; (b) moment-rotation at ground level; (c) bending
moment profile

especially at shallow depths. This may be caused by the under- Determination of Ultimate Soil Resistance
estimation of the α value in Fig. 9(b) and the unrealistic value of
This study adopts the method proposed by Reese et al. (1974) to
K 0 for overconsolidated dense sand.
calculate the ultimate soil resistance in the lateral direction
[Eq. (11)]. The values of the input parameters α and K 0 need
Proposed Design to be determined by referring to the specified ground condition,
which is usually ignored by other studies; it should not take de-
Based on the field lateral loading test results and the previous fault values α ¼ ϕ=2 and K 0 ¼ 0.4 for different ground condi-
analysis, the most commonly used API p-y model is adopted to be tions (API 2005), especially for overconsolidated dense sand
modified for the design of monopiles driven in dense sand deposits. deposits
Normalized soil reaction, p / D2
Normalized soil reaction, p / D2

250 80

200 1D,Measured 2D,Measured


3D,Measured 6D,Measured 60
1D,API 2D,API
150 3D,API 6D,API
40
100
20
50

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Normalized displacement, y/D (%) Normalized displacement, y/D (%)
(a) (b)

Fig. 14. Comparison between measured and API recommended p-y curves at various depth

© ASCE 04017026-9 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., -1--1


100 300
1D,Measured
1D,API

Normalized soil reaction,


250

Normalized soil reaction,


80 1D,This study
200
60

p /( D2)
p /( D2)
150
40
100
6D,Measured
20 6D,API
50 6D,This study

0 0
0 2 4 6 8 0 2 4 6 8
(a) Normalized displacement, y/D (%) (b) Normalized displacement, y/D (%)

Fig. 15. p-y curves comparison between measured, API 2005, and proposed by this study: (a) depth z ¼ 1D; (b) depth z ¼ 6D
Downloaded from ascelibrary.org by University of Newcastle on 03/12/17. Copyright ASCE. For personal use only; all rights reserved.

8 " #9
> K 0 z tan ϕ sin β tan β > deposit (API 2005; Klinkvort and Hededal 2014), i.e., m ¼ 1 in
< γz tanðβ−ϕÞ cos α þ tanðβ−ϕÞ ðD þ z tan β tan αÞ >
> = Eq. (12). For the overconsolidated dense sand deposits discussed
put ¼ min þK z tan βðtan ϕ sin β − tan αÞ − K D
>
>
0 a >
> in this study, the derived p-y curves show that the value of the ini-
: ;
γzD½K a ðtan8 β − 1Þ þ K 0 tan ϕtan4 β tial stiffness for each depth increases insignificantly with depth
(Fig. 14). To determine the distribution of initial stiffness with
ð11Þ
depth for this overconsolidated dense sand site, a power function
where α = angle defining the geometry of the wedge is fitted to the in situ measured profile of qc with a depth of up to
[Fig. 9(b)] as a function of soil density; K 0 = coefficient of earth 6 m, see Fig. 2(a), found that 10.9ðz=zref Þ0.35 fits well, where
pressure at rest; β ¼ 45° þ ϕ=2; and K a ¼ tan2 ð45° − ϕ=2Þ. zref ¼ 1 m. Therefore, the value of m in Eq. (14) is recommended
to be 0.35 by considering the general correlation between the soil
stiffness and cone tip resistance of CPT. This value agrees with
Determination of Initial Ground Stiffness current studies performed by Sorensen et al. (2010) and Kallehave
For the soil stiffness, the majority of existing research assumes et al. (2012), which suggest that the value of m ranges should be
that initial stiffness (K i ) increases linearly with depth (z) for sand from 0.3 to 0.7; Sorensen et al. (2010) made this recommendation

Bending moment, M (kN.m)


120 0 20 40 60
0.0
100

80
0.5
60 PS2
PS2 API
Depth, z (m)

40 API This study

20 This study 1.0

0
0 5 10 15 20 25
Lateral displacement, y0 (mm) 1.5
(a)

50 2.0

40
(c)

30

20
PS2
API
10 This study

0
0 0.2 0.4 0.6 0.8 1
Rotation, (degree)
(b)

Fig. 16. Lateral response comparison of PS2 between measured, and calculated with models of API (2005) and proposed by this study: (a) load-
displacement at ground line; (b) moment-rotation at ground line; (c) bending moment profile corresponding to F0 ¼ 60 kN

© ASCE 04017026-10 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., -1--1


by finite-element analysis, Kallehave et al. (2012) made it accord- 5. The API model overestimates the initial stiffness of overconso-
ing to the full-scale measurements, and both studies focused on the lidated dense sand at deeper depths and underestimates the ul-
large diameter monopiles timate soil resistance at shallow depths. This may be attributed
to the initial stiffness as a linearly distribution with depth and
K i ¼ ni · ðz=zref Þm ð12Þ default values of α ¼ ϕ=2 and K 0 ¼ 0.4, which ignores the
specified ground condition.
With regard to the value of ni in Eq. (14), a number of research- 6. The p-y model recommended by the current API is refined for
ers (e.g., Georgiadis et al 1992) found that the value of the initial laterally loaded rigid monopiles driven in dense sand deposit, in
stiffness in sand deposit is heavily overestimated by API (2005), which α ¼ ϕ, actual K 0 of overconsolidated dense sand, and a
and the recommended values by Georgiadis et al. (1992) and power distribution of initial stiffness are recommended. The
Kim et al. (2004) agree well with those given by Terzaghi (1955). modified API p-y model offered much better prediction, espe-
Therefore, the constant ni with the magnitude of 23.4 MPa and cially for bending moment profiles, than that calculated with the
corresponding to z ¼ 1 m depth (i.e., z ≈ 3D) is recommended API p-y model, which was originated for long-flexible piles.
Downloaded from ascelibrary.org by University of Newcastle on 03/12/17. Copyright ASCE. For personal use only; all rights reserved.

for the overconsolidated dense sand deposit.

Acknowledgments
Validation of Proposed Design
To validate the proposed design model, a comparison of p-y curves Field tests discussed in this paper are part of the first author’s doc-
and pile response is made between those measured and calculated toral research. Great support given by Dr. Ken Gavin, supervisor of
with the API (2005) recommendation and the proposed design the first author, is highly appreciated. Further thanks also go to the
model above. Fig. 15 shows a comparison of p-y curves at depth colleagues and technical staff in University College Dublin
of 1D and 6D between API and the present study. It clearly dem- who provided invaluable help. The authors acknowledge the finan-
onstrated that the p-y curve proposed by this study compares well cial support from National Natural Science Foundation of China
with the measured p-y curve in the available range. The compari- (No. 41502273) and Program for Young Excellent Talents in Tongji
son of the measured lateral responses of PS2 with calculations by University (No. 2015KJ009).
using API and the proposed p-y model mentioned previously is
illustrated in Fig. 16, which demonstrates that the p-y model pro-
posed by this study for overconsolidated dense sand deposits is References
more reasonable.
Abdel-Rahman, K., and Achmus, M. (2005). “Finite-element modelling of
horizontally loaded monopile foundations for offshore wind energy
converters in Germany.” Int. Symp. on Frontiers in Offshore, Taylor
Conclusion & Francis, Australia, 391–396.
API (American Petroleum Institute). (2005). “Recommended practice for
A series of field lateral loading tests on two reduced scale monop- planning, designing and constructing fixed offshore platforms: Working
iles were conducted at a dedicated geotechnical research site with stress design.” American Petroleum Institute, Washington, DC.
dense siliceous sand deposits. The ground soil condition is uniform, Bouzid, D. A., et al. (2013). “Winkler springs (p-y curves) for pile design
overconsolidated, and dense, which has been confirmed by exten- from stress-strain of soils: FE assessment of scaling coefficients using
sive excavation. The heavily instrumented model monopiles had an the mobilized strength design concept.” Geomech. Eng., 5(5), 379–399.
Bowman, E. R. (1958). “Investigation of the lateral resistance to movement
outer diameter of 0.34 m, an embedded length of 2.2 m, and a
of a plate in cohesionless soil.” Master’s thesis, Univ. of Texas, Austin,
slenderness ratio of 6.5. Both test piles with identical geometries TX.
and soil conditions were loaded statically, and pile head response DNV. (2007). Design of offshore wind turbine structures, Det Norske
and bending moment with depth were recorded. According to the Veritas, Oslo, Norway.
measured response, p-y curves at various depth and the lateral Doherty, P., and Gavin, K. (2012). “Laterally loaded monopile design for
shear force-displacement were derived. Based on the test results offshore wind farms.” Proc., ICE–Energy, 165(1), 7–17.
and analysis the following conclusions can be drawn: Doherty, P., Li, W., Gavin, K., and Casey, B. (2012). “Field lateral load test
1. Pile heads of both PS1 and PS2 responded similarly and in an on monopile in dense sand.” 7th Int. Conf. on Offshore Site Investiga-
elastic-plastic fashion. The displacement needed to mobilize the tion and Geotechnics, Integrated Technologies—Present and Future,
ultimate load capacity of both test piles is approximately 17 mm Society for Underwater Technology, London, 459–464.
Gavin, K., Doherty, P., and Tolooiyan, A. (2014). “Field investigation of
(0.05D) at ground level.
the axial resistance of helical piles in dense sand.” Can. Geotechn. J.,
2. As load increases, the numbers and size of the cracks developed 51(11), 1343–1354.
in front of the pile increased and demonstrated a failure Gavin, K. G., and O’Kelly, B. C. (2007). “Effect of friction fatigue on pile
model with the wedge shape. A further analysis on the distribu- capacity in dense sand.” J. Geotech. Geoenviron. Eng., 10.1061
tion of cracks shown on the ground surface showed that the va- /(ASCE)1090-0241(2007)133:1(63), 63–71.
lue of angle α (defining the geometry of the wedge) is equal to Georgiadis, M., Anagnostopoulos, C., and Saflekou, S. (1992). “Centrifu-
the peak friction angle of soil in such an overconsolidated gal testing of laterally loaded piles in sand.” Can. Geotechn. J., 29(2),
dense sand. 208–216.
3. For laterally loaded rigid monopiles in dense sand deposit, Hetenyi, M. (1946). “Beams on elastic foundation.” Univ. of Michigan,
the depth of rotation point varies in a narrow range of Ann Arbor, MI.
Jardine, R., Puech, A., and Andersen, K. H. (2012). “Cyclic loading
ð0.73−0.85ÞLem , and the smaller rotation depth for the higher
of offshore piles: Potential effects and practical design.” 7th Int.
load and for piles experiencing larger number of load cycles. Conf. on Offshore Site Investigation and Geotechnics: Integrated
4. The lateral shear force exists at the pile tip. This lateral shear Geotechnologies—Present and Future, Society for Underwater Tech-
force may play an important role for the analysis or design nology, London, 59–97.
of rigid monopiles. Since limited data is available in this study, Kallehave, D., Thilsted, C. L., and Liingaard, M. (2012). “Modification
a further study is warranted. of the API p-y formulation of initial stiffness of sand.” 7th Int. Conf.

© ASCE 04017026-11 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., -1--1


on Offshore Site Investigation and Geotechnics: Integrated Geotech- Murchinson, J. M., and O’Neill, M. W. (1984). “Evaluation of p-y relation-
nologies—Present and Future, Society for Underwater Technology, ships in cohesionless soils.” Analysis and design of pile foundations,
London, 465–472. ASCE, Reston, VA.
Kim, B. T., Kim, N.-K., Lee, W. J., and Kim, Y. S. (2004). “Experimental Nip, D. C. N., and Ng, C. W. W. (2005). “Back-analysis of laterally loaded
load-transfer curves of laterally loaded piles in Nak Dong River sand.” bored piles.” Proc., ICE: Geotechnical Eng., 158(2), 63–73.
J. Geotech. Geoenviron. Eng., 10.1061/(ASCE)1090-0241(2004)130:4 Peralta, P. (2010). “Investigation on the behavior of large diameter piles
(416), 416–425. under long-term lateral cyclic loading in cohesionless soil.” Leibniz
Kirwan, L. (2015). “Investigation into ageing mechanisms for axially Univ., Hannover, Germany.
loaded piles driven in sand.” Ph.D. thesis, Univ. College Dublin, Poulos, H. G., and Hull, T. S. (1989). “Role of analytical geomechanics in
foundation engineering.” Foundation engineering: Current principles
Dublin, Ireland.
and practices, ASCE, New York, 1578–1606.
Klinkvort, R. T., and Hededal, O. (2014). “Effect of load eccentricity
Reese, L. C., Cox, W. R., and Koop, F. D. (1974). “Analysis of laterally
and stress level on monopile support for offshore wind turbines.”
loaded piles in sand.” Proc., 6th Annual Offshore Technology Conf.,
Can. Geotech. J., 51(9), 966–974. Offshore Technology Conference, Dallas, 473–485.
Leblanc, C., Houlsby, G. T., and Byrne, B. W. (2010). “Response of stiff Reese, L. C., and Van Impe, W. F. (2001). Single piles and pile groups
Downloaded from ascelibrary.org by University of Newcastle on 03/12/17. Copyright ASCE. For personal use only; all rights reserved.

piles in sand to long-term cyclic lateral loading.” Geotechnique, 60(2), under lateral loading, Taylor & Francis, London.
79–90. Shadlou, M., and Bhattacharya, S. (2014). “Dynamic stiffness of pile in a
Lehane, B. M. (1992). “Experimental investigations of pile behaviour using layered elastic continuum.” Geotechnique, 64(4), 303–319.
instrumented field piles.” Ph.D. thesis, Imperial College, London. Sorensen, S. P. H., Ibsen, L. B., and Augustesen, A. H. (2010). “Effects of
Lesny, K., and Wiemann, J. (2006). “Finite-element modelling of large diameter on initial stiffness of p-y curves for large-diameter piles in
diameter monopiles for offshore wind energy converters.” GeoCongress sand.” 7th European Conf. on Numerical Methods in Geotechnical
2006, ASCE, Reston, VA, 1–6. Engineering, Taylor & Francis Group, London, 907–912.
Li, W., Igoe, D., and Gavin, K. (2014). “Evaluation of CPT-based p-y Terzaghi, K. (1955). “Evaluation of coefficients of subgrade reaction.”
models for laterally loaded piles in siliceous sand.” ICE-Geotech. Lett., Geotechnique, 5(4), 297–326.
4(2), 110–117. Tolooiyan, A., and Gavin, K. (2011). “Modelling the cone penetration
Li, W., Igoe, D., and Gavin, K. (2015). “Field tests to investigate the cyclic test in sand using cavity expansion and arbitrary Lagrangian eulerian
response of monopiles in sand.” Proc., ICE-Geotech. Eng., 168(5), finite-element methods.” Comput. Geotech., 38(4), 482–490.
407–421. Wesselink, B. D., Murff, J. D., Randolph, M. F., Nunez, I. L., and
Matlock, H. (1970). “Correlation for design of laterally loaded piles in soft Hyden, A. M. (1988). “Analysis of centrifuge model test data from lat-
clay.” Second Annual Offshore Technology Conf., Offshore Technology erally loaded piles in calcareous sand.” Proc., 1st Int. Conf. on Engineering
for Calcareous Sediments, A.A. Balkema, Rotterdam, Netherlands, 261–
Conference, Dallas, 577–594.
270.
Matlock, H., and Ripperger, E. A. (1956). “Procedures and instrumentation
Yan, L., and Byrne, P. M. (1992). “Lateral pile response to monotonic
for tests on a laterally loaded pile.” Proc., 8th Texas Conf. on Soil
pile head loading.” Can. Geotech. J., 29(6), 955–970.
Mechanics and Foundation Engineering, Bureau of Engineering Re- Yang, K., and Liang, R. (2007). “Methods for deriving p-y curves from
search, Univ. of Texas, Austin, TX. instrumented lateral load tests.” Geotech. Test. J., 30(1), 31–38.
McClelland, B., and Focht, J. A., Jr (1956). “Soil modulus for laterally Yoshida, I., and Yoshinaka, R. (1972). “A method to estimate modulus of
loaded piles.” J. Soil Mech. Found. Div., 82, 22. horizontal subgrade reaction for pile.” Soils Found., 12(3), 1–17.

© ASCE 04017026-12 J. Geotech. Geoenviron. Eng.

J. Geotech. Geoenviron. Eng., -1--1

You might also like