Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Marine Structures 69 (2020) 102684

Contents lists available at ScienceDirect

Marine Structures
journal homepage: http://www.elsevier.com/locate/marstruc

Field experiments on laterally loaded piles for an offshore


wind farm
Dong-sheng Xu a, Xue-yong Xu b, Wei Li c, Behzad Fatahi d, *
a
School of Civil Engineering and Architecture, Wuhan University of Technology, 122 Luoshi Rd., Wuhan 430070, PR China
b
Da Cheng Ke Chuang Foundation Construction Co.,Ltd, 297 HuaiHai Rd., Wuhan 430070, PR China
c
PowerChina Huadong Engineering Corporation Limited, Hangzhou, PR China
d
School of Civil and Environmental Engineering, University of Technology Sydney (UTS), Sydney, Australia

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

Keywords: Pile foundations are widely used to support offshore wind turbines due to their cost effectiveness
Offshore wind farm and rapid constructions. Offshore piles must be designed with enough capacity to withstand
Pile overturning moments caused by wind turbines and other environmental factors such as wave
Full-scale experiment
excitations and extreme winds. In this study, a full-scale field experimental test is undertaken to
Lateral deflection
determine the pile behaviour under various lateral loading conditions. A distributed fiber optic
sensing technology is used to measure strains along two instrumented piles. The bending mo­
ments and lateral deflections are calculated from distributed fiber optic sensors, and then ana­
lysed with the various p-y methods. Field measurements indicated that for two offshore piles ZK01
and ZK28 with diameter of 2 m and length of 71.5 m and 77.5 m, the maximum lateral move­
ments under a given lateral load of 800 kN were 369.1 mm and 351.7 mm, respectively. The
maximum bending moment occurred at 6.5 m and 5.5 m below seabed level with the corre­
sponding depth of 12.15D and 11.95D for pile ZK01 and ZK28, respectively. The position of “zero
crossing” of soil resistance for two instrumented piles is almost the same, even though the piles
have different lengths. The lateral deflections and bending moments of the two instrumented piles
are predicted by the API and hyperbolic method, which indicates that the hyperbolic method
yields larger prediction errors than the API method. A modified p-y approach is then proposed for
more reliable predictions when compared with field measurements.

1. Introduction

Offshore wind energy is growing rapidly in China in recent decades for moving towards green development. The common foun­
dations for offshore wind energy turbines include monopiles, tripods, jacket piles, suction buckets and gravity-based foundations. Piles
are the most widely used foundations (i.e. 86% of those currently used) due to their cost effectiveness, simple structure, and rapid
construction [1,2]. However, the bearing capacity and lateral deformations of offshore piles must meet strict guidelines due to the
higher lateral loads and overturning moments as compared to that in land environments [3].
A wide range of analytical, numerical, and physical model tests for analyzing the behaviour of piles have been reported in the

* Corresponding author. School of Civil and Environmental Engineering, Faculty of Engineering and Information Technology, University of
Technology, Sydney (UTS), City Campus PO Box 123, Broadway NSW 2007, Australia.
E-mail addresses: dsxu@whut.edu.cn (D.-s. Xu), behzad.fatahi@uts.edu.au (B. Fatahi).

https://doi.org/10.1016/j.marstruc.2019.102684
Received 8 January 2019; Received in revised form 25 July 2019; Accepted 3 October 2019
Available online 12 October 2019
0951-8339/© 2019 Elsevier Ltd. All rights reserved.
D.-s. Xu et al. Marine Structures 69 (2020) 102684

literature; for example, Klinkvort and Hededal [4] carried out centrifuge tests on model piles and proposed a new soil-pile interaction
curve; Poulos and Davis [5] examined the length of a pile affected its behaviour. In addition, many laboratory tests and numerical
methods have been utilised to investigate pile behaviour [6–11]. Among various analysis methods, the most commonly used approach
to analysis pile behavior under lateral loading is to replace the soil with Winkler springs, similar to a beam on a Winkler Foundation,
which is known as the p-y method [12]. The p-y approach was proposed by Reese et al. [13] for sands and was verified by many field
tests of small diameter piles. This method is usually suit for piles with diameters of less than 2 m [2]. The p-y curve approach had
indicated some disparities, especially in offshore situations [14]. DNV [15] guidelines also mentioned that the p-y curve method was
not calibrated for piles with large diameters. Researchers such as Hokmabadi et al. [16] and Chatziioannou et al. [17] have discussed
these concerns, and then extended this method to offshore piles. Murchinson and O’Neill [18] proposed a tangent hyperbolic function
for the p-y method, while Georgiadis et al. [19] developed a power law function and hyperbolic function for the p-y approach. The
design codes of API [12] and DNV [15] suggest using the tangent hyperbolic function for the p-y design method. The shear behavior of
sand with various mixtures may also influence the p-y curve for lateral loaded pile [20]. The soil-structure interaction (SSI), unloading
and reloading conditions related to the p-y method were also considered in many studies, such as Thieken et al. [21,22] and Achmus
et al. [23]. In summary, there are limited studies of piles tested under full-scale field conditions even though a wide range of analytical,
numerical, and physical model tests have been reported in the literature. A full-scale field test is a direct method of understanding the
nonlinear behaviour of piles in offshore conditions.
Fiber optic sensing technology now provides an effective approach to understand detailed mechanical behaviour of offshore piles.
One of the most commonly used sensing methods in civil engineering is the Brillouin scatter sensing technology, which is called the
Brillouin Optical Time Domain Analyzer (BOTDA) or Brillouin Optical Time Domain Reflectometer (BOTDR). This sensing technology
can measure full distributed strains with single mode fiber optic sensors (FOS). There are many instances where this sensing technology
has successfully been verified its effectiveness and efficiency in civil engineering [24–28]. In this study, the BOTDA sensing technology
was used in full scale field tests of two instrumented piles to investigate the lateral deformation of offshore piles. The variations in
strains along the piles were measured by optical sensors. The field instrumentations and test data were then analysed to: (1) establish a
benchmark study using BOTDA sensing technology of offshore piles in an offshore wind farm, (2) interpret offshore pile behavior under
different lateral loading conditions based on field test data; and (3) evaluate the p-y method based on field full-scale tests and then
propose an improved prediction approach.

2. Geological conditions and description of the site

2.1. Site location and geological conditions

The test site was located at the Xiangshui offshore wind farm in the Yellow Sea in China, as shown in Fig. 1. The offshore wind farm
was 10 km away from the coastline. It had a total area of 34.7 km2 with the dimension of 13.4 km in length and 2.6 km in width. The
whole offshore wind farm contains 50 wind turbines with the total capacity of 200 MW. The wind turbines have a diameter of 121 m
and the hub height of 85 m. The pile and elevated pile-cap foundation were used in this project. This offshore wind farm was con­
structed in an area where quaternary sediments containing mixed materials including silt, marine clay, silty sand and clay. The water
depth in the construction site varied from 8 m to 12 m. The highest, lowest and average tidal wave in this project site was 2.43 m,
2.18 m and 1.36 m, respectively. The field investigation was carried out by two integrated exploration platforms with offshore
drilling rig and C-Nav3050 GPS system. The field and laboratory investigation presented in this study were conducted in 2015 followed
by a detailed investigation in 2016.

2.2. Ground conditions

Ground stratigraphy was established from two boreholes located at the centre of the instrumented piles. Fig. 2 shows the borehole

Fig. 1. Locations of the study area.

2
D.-s. Xu et al. Marine Structures 69 (2020) 102684

log and ground conditions in the test field. In this testing site, five main soil deposits were detected, including marine deposit, soft clay,
silty sand, fine sand, and silty clay. The seabed is covered by 1.5–2 m of marine deposit and 24–26 m of grey brown soft clay with a thin
layer of silt, followed by 13–15 m of fine sand with occasional seashell gravel and humus. Beneath that layer, 20–22 m of firm to very
stiff silty sand was found with occasional shellfish gravel. The site investigation did not detect any bedrock. The standard penetration
test (SPT) was also conducted at the sampling borehole. The blow count N of SPT is plotted against the depth in Fig. 2. The N values
appear to increase with depth. Soil samples were also taken at different depths, thus the undrained shear strength cu was obtained and
is shown in Fig. 2. The SPT and cu values indicate shear strength of soil increased with depth. Additional in situ cone penetration tests
(CPT) were carried out and the results are shown in Fig. 2. Further laboratory test results on soft clay, fine sand, silty sand, and silty clay
are summarised in Table 1.

3. Investigation and testing program

3.1. Test layout

Fig. 3 shows the layout of field tests in this study. The instrumented piles were surrounded by 6 reaction piles (i.e. M1 - M6) and 2
reference piles (i.e. J1 and J2). The reference piles were used to install displacement sensors for settlement measurements of loaded
piles. After the loading tests, the reference piles were used as offshore foundations. The parameters of two instrumented piles are
presented in Table 2. The piles ZK01 and ZK28 are open-ended with 2.0 m in diameter and 30 mm in thickness. The total length of piles
ZK01 and ZK28 are 71.5 m and 77.5 m, respectively. The reaction piles M1 - M6 with diameter of 1.8 m and length of 74 m were used to
provide reaction forces during the loading tests. The instrumented piles ZK01 and ZK28 were installed on October 24 and November
15, 2014, respectively. The 6 reaction piles surrounding each instrumented pile were installed two weeks after the test pile in order to
diminish the effect of installation. The distance between reaction piles was 4.3 m which was 2.4 times of the pile diameter. The distance
between reaction pile and the instrumented pile was 5.5 m (centre to centre) which was 3.05 times of the reaction pile diameter. It is
expected that the impacts of reaction piles on the performance of the test pile would be minor considering the Saint Venant’s principle.
In addition, a field test was conducted by Brown et al. [9] in which the diameter of the test pile was 600 mm and the distance between
reaction piles and the instrumented pile was 2.1 m (i.e. 3.5 times of the diameter). The results by Brown et al. [9] showed no evident
interference from the reaction piles. Thus, in this study, it is expected that the reaction piles would have minor impacts on the test piles.

3.2. Pile installations

Two instrumented piles (i.e. ZK01 and ZK28) were installed using the impact driving with a diesel pile hammer. The maximum
impact forces generated by the hammer varied from 17.5 MN to 24.7 MN. The driving records during installation are shown in Fig. 4.
The records indicate that the soil resistances to driving increased significantly when the pile tip entered the silty sand layer. A pair of
strain gauges was installed below the pile head to measure the driving forces during installation; the maximum driving force for ZK01

Undrained shear Cone resistance


Test pile SPT N (blows/300 mm) strenght C (kPa) q /MPa
Sea water 0 20 40 60 80 0 25 50 75 100 0 5 10 15 20
0 0 0
Marine deposit
11.07m

Alluvium 20 20 20

36.72m
Depth below seabed level (m)

Fine sand 40 40 40

50.62m

Silty sand
60 60 60

71.32m
Silty clay
85.82m 80 80 80

Sleeve friction
f /kPa

100 100 100


0 50 100 150 200

Fig. 2. Ground conditions at test site in Yellow Sea, China. (For interpretation of the references to colour in this figure legend, the reader is referred
to the Web version of this article.)

3
D.-s. Xu et al. Marine Structures 69 (2020) 102684

Table 1
Properties for different soils from the testing site.
Property Soft clay Fine sand Silty sand Silty clay

Liquid limit, % 45.0 29.0 27.9 37.6


Plastic limit, % 18.6 20.8 18.6 22.8
Plasticity index, % 14.1 8.2 9.3 14.8
Moisture content, % 43.9 23.9 24.3 28.4
Dry density, g/cm3 1.22 1.62 1.64 1.53
Specific gravity of solids 2.73 2.68 2.72 2.73
Cc 0.98 0.07 0.11 0.33
e0 1.237 0.652 0.662 0.786

Fig. 3. Test piles: (a) photography of the test piles in field; (b) sketch of the pile arrangement.

4
D.-s. Xu et al. Marine Structures 69 (2020) 102684

Table 2
Summary of test results.
Pile No. ZK01 ZK28

Amplitude of pile head (meter above Principal Datum, mPD) þ7.8 þ7.9
Amplitude of lateral loading point (mPD) þ8.7 þ8.5
Pile length (m) 71.5 77.5
The maximum applied load, Hmax (kN) 800 800
The maximum deflection, ymax (mm) 467.7 436.5
The maximum bending moment, Mmax (kN⋅m) 16737 15363
Location of the maximum bending moment (mPD) 16.1 16.0

and ZK28 are 19.5 MN and 27.1 MN, respectively. It should be noted that the installation terminated when the driving record was
greater than 200 blows per meter. The penetration depths of ZK01 and ZK28 were 53.6 m and 59.1 m, respectively. The soil elevations
inside the piles ZK01 and ZK28 were 50.8 m and 56.6 m, respectively.

3.3. Instrumentations

The test piles were instrumented with a purpose designed single mode fiber optic sensor using BOTDA sensing technology to
measure distributed strains. The BOTDA sensing technology is based on light scattering inside the fiber. The backscattering signal is
detected when an optical pulse is launched into an optical fiber. According to Horiguchi et al. [29], the Brillouin frequency has a linear
relationship with the surrounding strains. Fig. 5 shows the measurement principle of the distributed strain sensing system.
By analyzing the reflected Brillouin signal, the frequency of Brillouin backscattered light is changed with the local temperatures
and strains. The relationship between Brillouin frequency and local strains and temperatures can be expressed as:
vB ðεÞ ¼ vB ð0Þ½1 þ Cε ε� (1)

vB ðTÞ ¼ vB ðT0 Þ½1 þ CT ðT T0 Þ� (2)

where v(ε) and v(T) represent Brillouin frequency shift at strain ε and temperature T, respectively, vB(0) and vB(T0) are the reference
frequencies, and CT and Cε are the temperature coefficient and strain coefficient, respectively.
A special type of single mode fiber optic sensor (FOS) was calibrated in the laboratory by applying incremental strains and then
used in this field test. A specific length of FOS fiber was fixed by two clamps and then various strains were applied. Fig. 6 shows the
calibration results. The frequency measured along the strained fiber and the unstrained fiber is shown in Fig. 6(a). The Brillouin

Blows/m
0 100 200 300 400

ZK01
Pile penetration (depth below seabed level, m)

30 ZK28

40

50

60

Fig. 4. Driving record for the two test piles (pile head is zeroed).

5
D.-s. Xu et al. Marine Structures 69 (2020) 102684

Fig. 5. Schematic of distributed strain sensing system.

frequency shift against the applied strains and temperatures are shown in Fig. 6(b) and (c), respectively. The results of this calibration
show that the Brillouin frequency shift has a linear relationship with the strains and temperatures. The strain coefficient Cε and
temperature coefficient CT are 471.5 MHz/% and 3.35 MHz/� C, respectively.
Fig. 7 shows the installing procedures of fiber optic sensors on the surface of the instrumented pile. The installation process could be
divided into three stages. In the first step, the surface of the pile was marked and polished to remove the iron rust so that the optical
fiber could glue properly; the fiber sensor was packed with well protection and then glued onto the surface of the piles. Secondly, the
designed fiber optic sensors were glued with epoxy resin and covered by ‘L’ shape steel. The fiber optic sensors were stretched before
being glued. Finally, the fiber optic sensors were connected to a fiber optic interrogator to record raw data. The initial data from fiber
optic sensor was extracted after the pile was installed. Before further testing, the piles were left untouched for two weeks after
installation to allow the pore water pressure dissipation.
Fig. 8 shows the layout of lateral loading tests with fiber optic sensors attached. Lateral loading tests were carried out with a re­
action system provided by 6 reaction piles. The loading system consisted of 10 hydraulic jacks with the maximum capacity of 5000 kN
and a 70 MPa oil pressure system. All the data was collected by a data-logger system. The lateral loads were controlled by load cells and
applied at the pile head in specific increments. In this study, the lateral loading tests were carried out one month after the axial loading
tests when it was thought that the excess pore water pressure would have been dissipated. Monotonic and cyclic loads were applied on
piles ZK01 and ZK28, respectively; while the maximum lateral load was 800 kN in the monotonic and cyclic loading tests. At each
loading stage, the lateral displacement and strains were measured by fiber optic sensors along the piles. Lateral displacements were
measured by the inclinometers, and the strains were measured by the BOTDA sensors. The tip of the pile was assumed to have zero
displacement and rotation due to the large embedded length of the pile, so the bending moment could be obtained from fiber optic
sensors.

3.4. Field test procedures

The field test was intended to evaluate their behaviour under monotonic and cyclic lateral loads. As mentioned earlier, the program
for testing piles against lateral loads consisted of reaction piles in two groups at their designed locations. Two sets of static lateral
loading tests were carried out on two independent piles ZK01 and ZK28. During the lateral loading tests, the pile head was in a free
rotation condition. The reaction forces were provided by 6 anchor piles, as shown in Fig. 3. A lateral load was applied onto the head of
the pile by hydraulic jacks. To avoid local damage during the lateral loading tests, the loading point on the head of the pile was
reinforced. The piles ZK01 and ZK28 were subjected to monotonic and cyclic loads, respectively. For pile ZK01, static lateral loads were
applied in increments of 250 kN until it reached 500 kN, then the increments were reduced to 100 kN until the total load reached
800 kN. For pile ZK28, static cyclic loads were applied in increments of 100 kN until a total load reached 800 kN. During each loading
step, the loads, lateral displacement at the pile head and strains using BOTDA strain sensors were recorded.

4. Test results and interpretation of data

4.1. Test results

The results of lateral loading tests are summarised in Table 2. The lateral deflections of instrumented piles increased gradually as
the applied lateral loads increased. The maximum lateral deflections occurred at the heads of piles, with the corresponding values for
ZK01 and ZK28 being 467.7 mm and 436.5 mm, respectively. The applied loads versus pile deflections at the heads for piles ZK01 and
ZK28 are plotted in Fig. 9. The measurements of the lateral loads versus lateral deflections show a nonlinear behaviour. The two piles
have similar hysteresis loop response due to the plastic characteristics of surrounding soils. However, the stiffness represented by the
relationship between lateral load and displacement was quite different. The pile ZK28 responded in a more rigid manner than ZK01
due to different loading conditions and pile lengths.

6
D.-s. Xu et al. Marine Structures 69 (2020) 102684

11.1
Applied strain 0.33%
Applied strain 0.28%
Applied strain 0.23%
11.0 Applied strain 0.19%
Applied strain 0.14%

Measured frequency (GHz)


Applied strain 0.09%
Applied strain 0.05%
10.9 Initial state

10.8

10.7

10.6
6.0 6.5 7.0 7.5
Distance along the fiber optic sensor (m)

(a)

11.00

Strain coefficient:
471.5 MHz/%
10.95
Brillouin frequency shift (GHz)

10.90

10.85

10.80

10.75
0.0 0.1 0.2 0.3 0.4
Strain (%)

(b)

Fig. 6. Calibration results of a distributed SMF with BOTDA sensing technology: (a) Measured Brillouin frequency shift of a strained SMF together
with the unstrained fiber; (b) Brillouin frequency shift with the applied strains; (c) Brillouin frequency shift with the surrounding tempera­
ture changes.

7
D.-s. Xu et al. Marine Structures 69 (2020) 102684

Fig. 7. Installation of distributed FOS: (a) special designed fiber optic sensor with two sets of sensors embedded in a thin layer of cooper steel; (b)
the surface of the pipe pile was polished to remove the iron rust for better glue of fiber sensors; (c) the designed fiber optic sensors was fixed on the
pipe pile surface, and (d) a ‘L’ shape steel was covered on the glued fiber optic sensor.

Load Monotonic

BOTDR
Sea leavel

Seabed level

y
z

Pile tip

Fig. 8. Layout of the lateral loading test.

4.2. Lateral deformation characteristics

The magnitude of bending moments at a given section of the pile can be obtained using the following equation:
Z
M ¼ σ ðxÞxdA (3)
A

where σ(x) is the normal stress in the cross section of the pile, and A is the area of the cross section. Considering the plane section of the
pile remains plane after loading, strains and stresses in the z-direction (axial direction of the pile) across the section of the pile can be
expressed as below:

du d2 y
εðxÞ ¼ ¼ x¼κ�x (4a)
dz dz2

8
D.-s. Xu et al. Marine Structures 69 (2020) 102684

1000
Pile ZK01
Pile ZK28
800

Lateral load (kN)


600

400

200

0
0 100 200 300 400 500 600
Lateral displacement (mm)

Fig. 9. Measured pile head responses of ZK01 and ZK28.

σðxÞ ¼ Ep εðxÞ ¼ Ep κ � x (4b)

where u is the displacement in the z direction, y is the deflection, κ is the curvature of the pile, Ep is the modulus of elasticity of pile
material. Thus, the bending moment can be derived as follows:
Z Z

M¼ Ep κ � x xdA ¼ Ep κ x2 dA ¼ Ep Ip κ (5)
A A

where Ip is the second moment of inertia of the pile and A is the cross-section area of pile. Based on the data measured at the pile
surface, the bending moment can be obtained by Eq. (5). According to the elastic beam theory for the laterally loaded pile, the reaction
of the soil (p(z)) on the laterally loaded pile can be expressed as below:

d4 y
pðzÞ ¼ Ep Ip (6)
dz4
Thus, with the distributed BOTDA sensors, soil resistance at any point along the pile can be determined using Eqs. (5) and (6). It
should be noted that the calculated soil resistances based on the assumption that the soil reaction is normal to the pile axis. Indeed, it is
vital to verify the accuracy of the assumption for the real case. In this study, pile deflections were measured by the inclinometer, and
therefore the p-y curves along the pile could be calculated from field measurements.
Fig. 10 shows the deflections of pile ZK01 and ZK28 under different lateral loads. As expected, the maximum deflection occurred at
the pile head and these deflections increased with the lateral loads. The depths of the zero-deflection points also tended to increase
with the lateral loads. After the lateral loads went beyond 500 kN, the depths of zero-deflection points increased slightly, and
deformation converged quickly. At a depth below 10 m (i.e., 10 m under the seabed level), lateral deflections were less than 14.3 mm
and 8.4 mm under the lateral load of 800 kN for ZK01 and ZK28, respectively, while the maximum deflections of the piles ZK01 and
ZK28 were 369.1 mm and 351.7 mm, respectively.
Fig. 11 shows the bending moments along the piles ZK01 and ZK28 under different lateral loads. At each loading stage, the bending
moments along the pile were obtained from distributed fiber optic sensors and then converted into curvatures. The optical fiber could
not be glued perfectly to the instrumented pile in the weld section due to the complex field conditions, thus it could introduce some
measurement errors at the welded section, and a sudden change in bending moment at this point could be observed in Fig. 11. The
results obtained from field tests indicate that the maximum bending moment occurred at 0.34H and 0.3H (H is the total length of the
pile) from the head of the ZK01 and ZK28 piles, respectively. The maximum bending moment occurred at a depth where lateral
deflection was small.
Fig. 12 shows the soil resistance along two instrumented piles under different lateral loading levels. The soil resistances at different
depths were calculated using Eqs. (4a) and (6). Note that the soil resistance increased as the lateral load increased. The profile of soil
resistance mainly depended on the soil strength; for instance, there was a sudden increase in soil resistance 10 m below the seabed
level, possibly due to the ground conditions shown in Fig. 2. Referring to Figs. 2 and 10 meters under the seabed is the first layer of sand
so the soil has a strong shear resistance to oppose pile deformation, and as Fig. 12(a) and (b) show, the position of “zero crossing” for
the ZK01 and ZK28 piles was almost the same at around 10 m below the seabed level. This could be due to the fact that the position of
“zero crossing” depends mainly on the pile-soil interaction characteristics, load eccentricity, pile stiffness, pile length and soil response.

9
D.-s. Xu et al. Marine Structures 69 (2020) 102684

Fig. 10. Measured pile deflections variation along the pile depth: (a) pile ZK01, (b) pile ZK28 (Depth: depth below the seabed level).

10
D.-s. Xu et al. Marine Structures 69 (2020) 102684

Fig. 11. Measured pile bending moments variation along the pile depth: (a) pile ZK01, (b) pile ZK28 (Depth: depth below the seabed level).

11
D.-s. Xu et al. Marine Structures 69 (2020) 102684

Fig. 12. Measured soil resistance variation along the pile depth: (a) pile ZK01, (b) pile ZK28 (Depth: depth below the seabed level).

12
D.-s. Xu et al. Marine Structures 69 (2020) 102684

5. Comparison between field measurements and predictions

5.1. The p-y methods currently used in practice

The response of a pile to lateral loads is usually analysed in practice by the API p-y method [12] which treats the soil-pile interaction

M0
P0

y
e
Sea water

z
Wedge failure

Shear failure
L

Fixed boundary

(a)

(b)
Fig. 13. Element of the lateral loaded monopile.

13
D.-s. Xu et al. Marine Structures 69 (2020) 102684

as a series of non-linear springs. The p-y method is widely used even though it was derived from a series of lateral loading tests with
small diameters. Referring to Reese et al. [13], the relationship between soil resistance (p) and lateral deflection of pile (y) can be
expressed as:
� �
kz
pðyÞ ¼ Apu tanh y (7)
Apu

where A is the coefficient to account for cyclic or static loading condition, A ¼ 0.9 for cyclic loading and A ¼ max (3–0.8x/D, 0.9) for
static loading, k is the initial modulus of subgrade reaction, pu is the ultimate soil resistance, z is the depth, y is the lateral displacement
and D is the diameter of the pile. The initial modulus of subgrade reaction could be determined as a function of the soil friction angle.
Georgiadis et al. [19] proposed a modified hyperbolic p-y curve based on the centrifuge tests as follows:
y
pðyÞ ¼ 1 y (8)
kz
þ Apu

5.2. Modified p-y method proposed in this study

In the p-y approach, two important factors will affect the prediction results, which are the initial modulus of the subgrade reaction
(k) and the ultimate soil resistance (pu). Many researchers have devoted significant efforts to improve their predictive accuracy by
modifying these two parameters; for instance Kallehave et al. [30] proposed a modified formulation of the initial stiffness for the p-y
approach as follows:
� �m � �0:5
z D
E ¼ kz ¼ kz0 (9)
z0 D0

where m was proposed to be equal to 0.6 based on the calibration results from a benchmark study, k is the initial subgrade reaction
modulus, z0 is a reference depth, D and D0 are the diameter and a reference diameter, respectively.
The soil surrounding the pile would be sheared due to the lateral movement. Thus, the maximum soil resistance is called the ul­
timate soil resistance (pu). The distribution of pu has been studied by many researchers, and can be classified as three categories [31]:
(a) the distribution of pu was obtained based on in-situ field tests, such as Broms [32] and Barton [33]; (b) the distribution of pu was
derived by a combination of in-situ field tests and theoretical analysis such as Reese [13] and Matlock [34]; and (c) the distribution of

Fig. 14. Flowchart for analysis of monopile behavior with the modified p-y method.

14
D.-s. Xu et al. Marine Structures 69 (2020) 102684

Lateral displacement, y (mm) Bending moment, M (kN.m)


0 20 40 60 80 100 0 5000 10000 15000 20000
-20 -20

-10
-10
Depth below seabed level, z (m)

Depth below seabed level, z (m)


0
0

10
ZK01
10 ZK01
250 kN, Field results in this study
600 kN, Field results in this study 250 kN, Field results in this study
20
800 kN, Field results in this study 600 kN, Field results in this study
250 kN, API [12] 800 kN, Field results in this study
600 kN, API [12] 250 kN, API [12]
20
800 kN, API [12] 600 kN, API [12]
30 800 kN, API [12]
250 kN, Hyperbolical method [19]
600 kN, Hyperbolical method [19] 250 kN, Hyperbolical method [19]
800 kN, Hyperbolical method [19] 600 kN, Hyperbolical method [19]
30 800 kN, Hyperbolical method [19]
40

(a) (b)

Lateral displacement, y (mm) Bending moment, M (kN.m)


0 20 40 60 80 100 0 5000 10000 15000 20000
-20 -20

-10
-10
Depth below seabed level, z (m)

0
Depth below seabed level, z (m)

10
ZK28
10 200 kN, Field results in this study
ZK28
600 kN, Field results in this study 200 kN, Field results in this study
20 600 kN, Field results in this study
800 kN, Field results in this study
200 kN, API [12] 800 kN, Field results in this study
600 kN, API [12] 200 kN, API [12]
20 600 kN, API [12]
800 kN, API [12] 30
200 kN, Hyperbolical method [19] 800 kN, API [12]
600 kN, Hyperbolical method [19] 200 kN, Hyperbolical method [19]
800 kN, Hyperbolical method [19] 600 kN, Hyperbolical method [19]
800 kN, Hyperbolical method [19]
30 40

(c) (d)
Fig. 15. Comparisons of the API method and the hyperbolical method: (a) lateral displacement of monopile ZK01; (b) bending moment of monopile
ZK01; (c) lateral displacement of monopile ZK28; (d) bending moment of monopile ZK28.

15
D.-s. Xu et al. Marine Structures 69 (2020) 102684

Lateral displacement, y (mm) Bending moment, M (kN.m)


0 10 20 30 40 50 60 70 80 90 0 5000 10000 15000 20000
-20 -20

-10
-10

Depth below seabed level, z (m)


Depth below seabed level, z (m)

0
0

10
ZK01
10 ZK01
250 kN, Field results in this study 250 kN, Field results in this study
600 kN, Field results in this study 20 600 kN, Field results in this study
800 kN, Field results in this study 800 kN, Field results in this study
250 kN, API [12] 250 kN, API [12]
20 600 kN, API [12] 600 kN, API [12]
800 kN, API [12] 30 800 kN, API [12]
250 kN, Proposed method 250 kN, Proposed method
600 kN, Proposed method 600 kN, Proposed method
800 kN, Proposed method 800 kN, Proposed method
30 40

(a) (b)
Lateral displacement, y (mm) Bending moment, M (kN.m)
0 20 40 60 80 100 0 5000 10000 15000 20000
-20 -20

-10
-10
Depth below seabed level, z (m)
Depth below seabed level, z (m)

10
ZK28
10 ZK28
200 kN, Field results in this study
200 kN, Field results in this study
600 kN, Field results in this study 20
600 kN, Field results in this study
800 kN, Field results in this study
800 kN, Field results in this study
200 kN, API [12]
200 kN, API [12]
600 kN, API [12]
20 600 kN, API [12]
800 kN, API [12] 30
800 kN, API [12]
200 kN, Proposed method
200 kN, Proposed method
600 kN, Proposed method 600 kN, Proposed method
800 kN, Proposed method 800 kN, Proposed method
30 40

(c) (d)
Fig. 16. Measured and derived profiles of (a) lateral displacement of monopile ZK01; (b) bending moment of monopile ZK01; (c) lateral
displacement of monopile ZK28; (d) bending moment of monopile ZK28.

16
D.-s. Xu et al. Marine Structures 69 (2020) 102684

pu was derived by theoretical analysis, for example Gluskov [35]. As shown in Fig. 13(a), the failure of soil surrounding the pile can be
divided in two layers; in the surface layer the soil fails in a wedge shape due to the small overburden stress, but in the deeper layer the
soil fails following a spoiler failure model due to the large overburden stress. Guo [31] make important progress to propose a unified
distribution of pu for considering both clay and sandy soils at the same time. Based on these studies, a modified distribution of pu with
depth is proposed in this study as follows:
� 0 2 n
γ � D � ðα0 þ zÞn ​ ​ ​ for ​ clay ​ soil ​ ​
pu ¼ (10)
βKp 2 � cu � D1 n � zn ​ ​ ​ ​ for ​ sandy ​ soil

where Kp ¼ tan2 ð45 þ φ =2Þ, ϕ is the friction angle, β is the coefficient (i.e. β ¼ 0.5–3), D and z are the diameter of the pile and the depth
of point of interest below the surface, respectively, n is a constant (i.e. n ¼ 0.7–0.8 for clay soil; n ¼ 1.0–1.8 for sand soil), α0 is the
parameter to represent the ultimate soil resistance at the surface which is equal to 0 for cohesionless soils. The value of βK2p can be
determined by the ultimate soil resistance at the surface. Thus, the modified p-y approach was constructed based on Eqs. (7), (9) and
(10).
In order to evaluate the pile behaviour, an element of the laterally loaded pile is considered as shown in Fig. 13(b). The pile was
considered as a linear elastic material with modulus (E) and second order moment of the cross section (I), respectively. Based on the
equilibriums of lateral forces and moments and considering the pile was laterally supported by surrounding soil with lateral
displacement y, the equilibrium equation can be derived and then it is solved using the finite difference method. The pile is divided into
nþ2 elements and therefore, the equilibrium equation can be rewritten as the following form:

yiþ2 4yiþ1 þ 6 þ δi h4 yi 4yi 1 þ yi 2 ¼ 0 (11a)
p ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
δi ¼ 4
ki D=ð4EIÞ; i ¼ 0; 1; 2; …; n (11b)

where h is the differential interval between two between element i and iþ1, ki is the subgrade modulus stiffness of the element i which
is corresponding to kz in Eq. (7). The boundary condition can be regarded as a point load (i.e. P0 in Fig. 13(a)) and bending moment (i.
e. M0 in Fig. 13(a)) applied at the head of the monopile. The bottom of the monopile was assumed fixed with no movement and
rotations. Thus, the pile deflection, bending moment, shear force and soil reaction can be calculated based on Eqs. (9)–(11) and
boundary conditions. Fig. 14 shows the step by step calculation procedures. In summary, (a) the monopile is first divided into n el­
ements; (b) the soil parameters, initial modulus of the subgrade reaction K (i.e. Eq. (9)) and boundary conditions are used to obtain the
matrix of lateral displacements [U]1; and (c) the soil resistance p is obtained based on the calculated lateral displacement [U]1, ultimate
soil resistance (i.e. Eq. (10)) and p-y curve (i.e. Eq. (7)). Then, the modulus of subgrade reaction is updated to a new matrix K, and a
new matrix of lateral displacements [U]2 corresponding to the updated modulus of subgrade K obtained from step (b); (d) steps (b) and
(c) are repeated until predicted [U]1 and [U]2 converge (i.e. errors less than 10 4); and (e) the bending moment, shear force and soil
reactions are calculated based on the determined lateral displacements.
Fig. 15 shows the comparisons between field measurements and calculated results from the API method [12] and the hyperbolic
method proposed by Georgiadis et al. [19]. Although the predictions of the API method and the hyperbolic method were generally in
good agreement with the field measurements, both approaches predicted larger lateral deflections at the pile head in comparison to the
field measurements. In addition, as shown in Fig. 15, there are some disparities between predicted bending moments and measured
results for the depth above 10 m. However, referring to Fig. 15(b), it can be observed that the API method resulted in slightly better
predictions in comparison to the hyperbolic method in this case study.
The modified p-y approach was proposed in this study and explained above in details which was used to predict the laterally
performance of loaded piles. The calculations are compared with field testing results and the API predictions which are shown in
Fig. 16. The estimated pile deflections and bending moments of the pile estimated by the API method was also plotted in Fig. 16 for
comparison purposes. In addition, The prediction errors for maximum lateral deflection and bending moment based on the API
method, hyperbolic method and the proposed method in this study are summarised in Table 3. The total average errors are 4.26%,
7.21% and 2.55% for the API method, hyperbolic method and the proposed method in this study, respectively. It can be also observed
that the API method has errors of 17.1% and 10.3% corresponding to the maximum lateral deflections for piles ZK01 and ZK28
respectively. The proposed modified p-y approach resulted in more reliable predictions and less errors (i.e. the corresponding errors of
3.9% and 8.0%, respectively) in comparison to other two methods. However, it should be also noted that some differences were
observed between the predicted and measured bending moments, which could be attributed to various reasons, such as the uncertainly
in soil-structure interaction mechanism along the pile, and load and pile eccentricity during the testing. However, referring to Table 3,
it is evident that the average bending moment prediction errors between the proposed method and the measured results were 0.1% and
4.0% for piles ZK01 and ZK28, respectively. By considering the complex conditions in the field, the predictions are in good agreement
with the measurements.
In summary, with the advanced distributed fiber optic sensors, the distributed strains, axial forces and lateral deflections of the pile
can be examined carefully. Thus, the distributed or quasi-distributed bending moments and soil resistances along instrumented piles
can be obtained which could be useful for theoretical analysis of soil-pile interaction. The API approach was verified by field full-scale
experimental tests which revealed that the API method could give acceptable prediction results. Although this method was quite
simple and widely adopted in engineering practices, the predicted results of bending moments could be further improved. In this study,
the modified p-y approach was taken into consideration of failure patterns of soils surrounding the pile by using a modified distribution

17
D.-s. Xu et al. Marine Structures 69 (2020) 102684

Table 3
Comparison between field measurements and predictions.
Pile Type Lateral Test API method [12] Hyperbolic method This study
No. load/kN results [19]

Calculated Error Calculated Error Calculated Error

zk01 Maximum lateral deflection along 250 16.5 21.7 31.5% 21.5 30.3% 18.4 11.5%
embedded length (mm) 600 58.2 64.5 10.8% 64.8 11.3% 58.3 0.2%
800 82.9 90.3 8.9% 91.3 10.1% 82.94 0.05%
Maximum bending moment (kN.m) 250 5525 5325 3.6% 5325 3.6% 5347 3.2%
600 14043 13549 3.5% 13540 3.6% 13366 4.8%
800 16737 17604 5.2% 18511 10.6% 18130 8.3%
zk28 Maximum lateral deflection along 250 15.5 13.3 14.2% 16 3.2% 14.2 8.4%
embedded length (mm) 600 70.3 68.7 2.3% 71.8 2.1% 70 0.4%
800 86.8 99.2 14.3% 103.4 19.1% 100.1 15.3%
Maximum bending moment (kN.m) 250 4448 3956 11.1% 3974 10.7% 3972 10.7%
600 12357 12553 1.6% 12738 3.1% 13036 5.5%
800 15363 17436 13.5% 17607 14.6% 17998 17.2%

of ultimate soil resistances. From the comparisons, it found that the modified approach has better predictions about the bending
moments.

6. Conclusions

A full-scale field experimental program carried out at an offshore wind farm in the Yellow sea of China aimed at examining the
behaviour of offshore piles in field conditions. The lateral deformation characteristics of two instrumented piles were analysed by field
tests. The results indicated that the maximum lateral movement of the steel tubular piles ZK01 (i.e. diameter of 2 m, length of 71.5 and
embedment depth of 53.9 m) and ZK28 (i.e. diameter of 2 m, length of 77.5 and embedment depth of 59.9 m) under a lateral load of
800 kN was 369.1 and 351.7 mm, respectively, and the maximum lateral deflection decreased by 17.4 mm as the pile length increased
6 m from ZK01 to ZK28. The distributions of bending moments and soil resistance along the pile were computed based on the strains
measured by the fiber optic sensors. The maximum bending moment occurred at 6.5 m and 5.5 m below seabed and the corresponding
depths were 12.15D and 11.95D for piles ZK01 and ZK28, respectively.
In addition, the lateral deflections and bending moments of the two piles were predicted by the API and other methods. The API
method indicated predictive errors of 17.1% and 10.3% corresponding to the maximum lateral deflections for piles ZK01 and ZK28
respectively. A modified p-y approach resulted in more reliable predictions and less error (i.e. the corresponding errors of 3.9% and
8.0%, respectively) in comparison to existing techniques available. The outcomes of this study can be used as references for design of
offshore piles, similar to what was used in the presented project.

Acknowledgement

The authors gratefully acknowledge the financial support provided by National Natural Science Foundation of China (project No.
41972271) and the Fundamental Research Funds for the Central Universities (WUT, project No. 193106001). Financial support from
Zhejiang Province Natural Science Foundation of China (LY14D020001) is also appreciated. The authors also acknowledge and
appreciate the anonymous reviewers for their comments and suggestions which enhanced the quality of this paper significantly.

Notation

A the cross section area of piles


CT temperature coefficient
Cε strain coefficient
D the diameter of the pile;
D0 a reference diameter
Ep the modulus of elasticity of pile material
ε strain
H the total length of the pile;
Ip the second moment of inertia of the pile;
Kp the passive earth pressure coefficient
k the initial modulus of subgrade reaction
M bending moment
M0 bending moment at pile head
m coefficient
n a constant

18
D.-s. Xu et al. Marine Structures 69 (2020) 102684

σ(x) the normal stress in the cross section of the pile;


P0 applied point load at pile head
p(z) soil resistance at depth of z
pu the ultimate soil resistance
T temperature
u the displacement in the z direction
v(ε) Brillouin frequency shift at strain ε
v(T) Brillouin frequency shift at temperature T
vB(0) reference Brillouin frequency at initial strain
vB(0) reference Brillouin frequency at initial temperature T0
y deflection
z0 a reference depth
α0 the parameter to represent the ultimate soil resistance at the surface
β a coefficient
κ the curvature of the pile;
δi coefficient

References

[1] Shadlou M, Bhattacharya S. Dynamic stiffness of monopiles supporting offshore wind turbine generators. Soil Dyn Earthq Eng 2016;88:15–32.
[2] Walsh C, WindEurope. Offshore wind in Europe - key trends and statistics 2018. WindEurope; 2019. p. 14–9.
[3] Gavin K, Igoe D, Doherty P. Piles for offshore wind turbines: a state-of-the-art review. P I Civil Eng-Geotec 2011;164:245–56.
[4] Klinkvort RT, Hededal O. Effect of load eccentricity and stress level on monopile support for offshore wind turbines. Can Geotech J 2014;51:966–74.
[5] Poulos HG, Davis EH. Pile foundation analysis and design. New York: John Wiley and Sons; 1980.
[6] Zhang YH, Andersen KH. Scaling of lateral pile p-y response in clay from laboratory stress-strain curves. Mar Struct 2017;53:124–35.
[7] Klar A, Randolph MF. Upper-bound and load-displacement solutions for laterally loaded piles in clays based on energy minimisation. Geotechnique 2008;58:
815–20.
[8] Zhang Y, Andersen KH, Tedesco G. Ultimate bearing capacity of laterally loaded piles in clay - some practical considerations. Mar Struct 2016;50:260–75.
[9] Brown MJ, Hyde AFL, Anderson WF. Analysis of a rapid load test on an instrumented bored pile in clay. Geotechnique 2006;56:627–38.
[10] Zhang Y, Andersen KH. Scaling of lateral pile p-y response in clay from laboratory stress-strain curves. Mar Struct 2017;53:124–35.
[11] Morato A, Sriramula S, Krishnan N, Nichols J. Ultimate loads and response analysis of a monopile supported offshore turbine using fully coupled simulation.
Renew Energy 2017;101:126–43.
[12] API RP 2A LRFD. Recommended practice for planning, designing and constructing fixed offshore platforms, load and resistance factor design. Technical report.
first ed. USA: API; 1993.
[13] Reese LC, Cox WR, Koop FD. Analysis of laterally loaded piles in sand. In: Proceedings of the sixth annual offshore technology conference, Houston, TX, USA
may 6-8; 1974.
[14] Lesny K, Paikowsky SG, Gurbuz A. Scale effects in lateral load response of large diameter monopiles. Contemporary Issues in Deep Foundations, vol. 158.
Geotechnical special publication; 2007. p. 1–10.
[15] DNV. Design of offshore wind turbine structures. Det Norske Veritas; 2014.
[16] Hokmabadi AS, Fakher A, Fatahi B. Full scale lateral behaviour of monopiles in granular marine soils. Mar Struct 2012;29:198–210.
[17] Chatziioannou K, Katsardi V, Koukouselis A, Mistakidis E. The effect of nonlinear wave-structure and soil-structure interactions in the design of an offshore
structure. Mar Struct 2017;52:126–52.
[18] Murchison JM, O’Neil MW. Evaluation of p-y relationships in cohesionless soils. In: Proceedings of the geotechnical engineering division, ASCE National
convention. Calif: San Francisco; Oct. 1-5; 1984. p. 174–92. Analysis and Design of Pile Foundations.
[19] Georgiadis M, Anagnostopoulos C, Saflekou S. Centrifugal testing of laterally loaded piles in sand. Can Geotech J 1992;29:208–16.
[20] Xu DS, Tang ZY, Zhang L. Interpretation of coarse effect in simple shear behavior of binary sand-gravel mixture by DEM with authentic particle shape. Constr
Build Mater 2019;195:292–304.
[21] Thieken K, Achmus M, Lemke K, Terceros M. Evaluation of p-y approaches for large diameter monopiles in sand. Int J Offshore Polar Eng 2015;2:134–44.
[22] Thieken K, Achmus M, Terceros M, Lemke K. Evaluation of p-y approaches for large-diameter piles in layered sand. Int J Offshore Polar Eng 2018;3:318–27.
[23] Achmus M, Thieken K, Saathoff J, Terceros M, Albiker J. Un- and reloading stiffness of monopile foundations in sand. Appl Ocean Res 2019;84:62–73.
[24] Zhu HH, She JK, Zhang CC, Shi B. Experimental study on pullout performance of sensing optical fibers in compacted sand. Measurement 2015;73:284–94.
[25] Xu DS, Liu HB, Luo WL. Evaluation of interface shear behavior of GFRP soil nails with a strain-transfer model and distributed fiber-optic sensors. Comput
Geotech 2018;95:180–90.
[26] Xu DS, Yin JH. Analysis of excavation induced stress distributions of GFRP anchors in a soil slope using distributed fiber optic sensors. Eng Geol 2016;213:
55–63.
[27] Zhu HH, Shi B, Zhang CC. FBG-based monitoring of geohazards: current status and trends. Sensors 2017;17:452.
[28] Hong CY, Zhang YF, Li GW, Zhang MX, Liu ZX. Recent progress of using Brillouin distributed fiber sensors for geotechnical health monitoring. Sens Actuators A-
Phys 2017;258:131–45.
[29] Horiguchi T, Kurashima T, Tateda M. Tensile strain dependence of Brillouin frequency shift in silica optical fibers. IEEE Photonics Technol Lett 1989;1:107–8.
[30] Kallehave D, Thilsted CL, Liingaard M. Modification of the API p-y formulation of initial stiffness of sand. In: 7th international conference: offshore site
investigation and geotechnics: integrated geotechnologies present and future, London; 2012. p. 465–72.
[31] Guo F, Lehane BM. Experimentally derived CPT-based p-y curves for soft clay. In: The 3rd international symposium on cone penetration testing. Las Vegas;
2014. p. 1021–8.
[32] Broms BB. Lateral resistance of piles in cohesionless soils. J Soil Mech Found Eng 1964;90:123–56.
[33] Barton YO. Laterally loaded piles in sand: centrifuge tests and finite element analysis. Ph.D Thesis. London: University of Cambridge; 1982.
[34] Matlock H. Correlations for design of laterally loaded piles in soft clay. In: 2nd offshore tech conf. Dallas; 1970. p. 577–94.
[35] Gluskov GN. Design of installations embedded in soil. Moscow: Stroiizdat; 1977.

19

You might also like