Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Hu, Q. et al. (2022). Géotechnique 72, No. 12, 1035–1050 [https://doi.org/10.1680/jgeot.20.

00002]

Lateral load response of large-diameter monopiles in sand


QIAN HU †, FEI HAN†‡, MONICA PREZZI†, RODRIGO SALGADO† and MINGHUA ZHAO 

Due to the increasing demand for renewable clean energy, there has been rapid growth in the development
of offshore wind energy resources in recent years. The majority of offshore wind turbines are founded on
single, large-diameter, open-ended pipe piles known as monopiles. The design of these piles is often done
using the p–y method, which models the soil simply as non-linear springs placed along the pile length and
assumes the pile to behave as a one-dimensional beam. To develop a design method that accounts for
three-dimensional pile–soil interaction and that is applicable to general conditions, a series of finite-
element (FE) analyses covering a wide range of pile dimensions, wall thicknesses, slenderness ratios, load
eccentricities, sand types and sand relative densities were performed using an advanced sand constitutive
model. In this paper, a set of equations is proposed that can be used to estimate the critical pile length,
Lcrit (the pile length beyond which the pile lateral capacity no longer increases), the lateral capacity, H,
and the lateral load–rotation curve for monopiles. The proposed method produces estimates of the lateral
capacity of monopiles that are in very close agreement with those from the FE analyses.

KEYWORDS: bearing capacity; finite-element modelling; foundations and soil–structure interaction;


monopile; offshore engineering; sands; wind turbine

INTRODUCTION Euler–Bernoulli or Timoshenko beam in contact with the


Offshore wind energy has gained popularity as a clean soil, which in turn is modelled as a series of non-linear
renewable energy resource because of the steady supply of springs. Each spring is defined by a relationship between the
fast wind in the open seas. The total global capacity of lateral soil resistance p per unit length of pile and the pile
installed offshore wind turbines has increased from 1 GW in deflection y at the depth where the spring is placed (Reese
2008 to 23 GW in 2018, and is expected to reach 100 GW by et al., 1975); hence the method designation as the p–y
2025 (GWEC, 2018). For reference, the US average elec- method. The method was empirically derived using data
tricity consumption rate was 477 GW in 2018 (US Energy obtained from a limited number of lateral load tests
Information Administration (EIA, 2019)). It is estimated that performed on long, slender piles (e.g. Cox et al., 1974;
the design, installation and construction of foundations of Reese et al., 1974). The pile soil analysis (PISA) research
offshore wind turbines costs about 20–30% of the total cost project (Byrne et al., 2015, 2019a, 2019b; Zdravković et al.,
of the wind turbine (Byrne et al., 2019a); thus, meaningful 2015, 2020; Burd et al., 2020a, 2020b; McAdam et al., 2020;
savings can be realised with optimised designs. From all the Taborda et al., 2020) examined ways to modify the p–y
foundation options that are available – gravity base, mono- method based on results of field load tests and three-
piles, jackets/tripods and floating foundations – monopiles, dimensional (3D) finite-element (FE) analyses. The modified
which are single open-ended pipe piles with diameters as large p–y method was developed for two specific soil profiles (stiff
as 10 m, are the most commonly used solution (Doherty & glacial clay till deposits in Cowden, and dense marine sand at
Gavin, 2012; Randolph & Gourvenec, 2017). The perform- the Dunkirk site). Additional FE analyses (in a second phase
ance of offshore wind turbines is largely dependent on the of the PISA project) have been performed to extend the
performance of the foundation elements, since a wind turbine application of the modified p–y method to a wider range of
is often supported on a single pile. Fig. 1 shows a schematic relative densities for Dunkirk sand (Byrne et al., 2019b).
representation of an offshore wind turbine supported on a In this paper, a series of 3D FE analyses of laterally loaded
monopile with a diameter B and an embedded length L. The monopiles was performed considering pile diameters ranging
horizontal wind load H acts through the rotor blades at the from 1 to 10 m, slenderness ratios L/B ranging from 3 to 20,
top of the turbine tower at a height h from the mudline, wall thickness-to-diameter ratios tw/B ranging from 1:100 to
resulting in a lateral load H and a moment equal to Hh 1:50, two types of sand (Toyoura sand and Ottawa sand) with
applied on the monopile at the mudline. The height h is also different particle morphologies, relative densities ranging from
known as the load eccentricity for the pile. 40 to 80% and load eccentricities ranging from 15 to 30 m.
The design of large-diameter offshore monopiles has Based on the results of these 3D numerical simulations, the
mostly relied on the ‘p–y method’ (API, 2014; DNV GL, authors propose design equations for monopiles subjected to
2016), which models the pile as a one-dimensional (1D) lateral loads that produce designs that are equivalent to those
produced by the 3D FE analyses. The proposed equations can
be used to obtain the complete lateral load–rotation curve for
a monopile and to estimate the lateral capacity of the pile
corresponding to any rotation at the mudline (within the
Manuscript received 30 August 2020; revised manuscript accepted rotation range considered in this paper). A web application
12 May 2021. Published online ahead of print 27 September 2021.
Discussion on this paper closes on 1 April 2023, for further details
was developed based on the results of the simulations for
see p. ii. convenient application of the proposed method.
 College of Civil Engineering, Hunan University, Changsha,
P. R. China.
† Lyles School of Civil Engineering, Purdue University, West FINITE-ELEMENT MODELLING
Lafayette, IN, USA. The load response of monopiles subjected to lateral loads
‡ University of New Hampshire, Durham, NH, USA. varies as a function of the diameter B, length L and wall

1035

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
1036 HU, HAN, PREZZI, SALGADO AND ZHAO
from calibration of the model against the results of elemental
tests (e.g. TXC, TXE and SS) performed on Ottawa sand and
Blade Toyoura sand. Table 2 provides the basic properties of these
two sands (Loukidis, 2006). As this model is not available in
H
Abaqus Explicit (Abaqus, 2014), the model was coded as a
user-defined material subroutine (Vumat) in Fortran (Han
et al., 2017). The steel pipe pile was modelled as a linear
elastic material with a Young’s modulus Ep = 200 GPa and a
Poisson’s ratio ν = 0·2, and the unit weight of the pile was
assumed to be the same as that of the sand.
Tower
h

Finite-element mesh and boundary conditions


Figure 2 shows the geometric dimensions of the FE model
and the mesh configuration used for the FE analysis of a pile
Mudline with B = 2 m, L = 10 m and h = 15 m. A total number of
(ground surface) 49 584 elements were used in this simulation. Owing to
L Pile geometric and loading symmetries, only half of the whole
B domain was modelled. Linear, eight-node, hexahedral
elements with reduced integration were used to model both
the pile and the soil domain. The impact of mesh size on the
Fig. 1. A typical monopile and the loading that it must carry FE analysis of laterally loaded piles was investigated by
performing three analyses of the same laterally loaded pile
(B = 2 m and L = 20 m with h = 15 m and DR = 80%) using
thickness tw of the pile; the sand type and its relative density three different meshes, with smallest element sizes of 1 cm,
DR; and load eccentricity h. As summarised in Table 1, a 4 cm and 10 cm. As Fig. 3 shows, the effect of mesh size is
wide range of values was considered for these parameters in negligible, at least for the pile rotation (and deflection) level
the FE analyses reported here. considered in the present study. Thus, the 4 cm mesh was
used in the analyses.

Constitutive model
When a pile is loaded laterally, sand elements located at Soil–pile interface
different positions with respect to the pile experience a variety Two approaches are commonly used to model the soil–pile
of loading paths, notably: approximately triaxial com- interface: (a) the contact pair approach, which defines a pair of
pression (TXC) and triaxial extension (TXE), respectively, master and slave surfaces and specifies the contact behaviour
ahead of and behind the pile near the ground surface; the between the two surfaces in the normal and tangential
opposite near the pile base; and approximately simple shear directions; and (b) the perfect contact approach, which
(SS) near the rotation point and below the base of short stiff assumes that no sliding occurs at the soil–pile interface. In
piles. It is essential to select a constitutive model that is able this second approach, the common nodes of the soil and pile
to accurately capture sand behaviour under all these loading are tied to each other with respect to all degrees of freedom
paths in order for the simulation results to be useful in (this approach is often used to model the pile–soil interface for
practice. The advanced two-surface-plasticity constitutive non-displacement piles (Salgado et al., 2017; Han et al.,
model developed by Loukidis & Salgado (2009) was used in 2019b)). To examine the effect of the interface modelling on
the FE analyses. The constitutive model follows the critical the laterally loaded pile problem, comparisons were made
state soil mechanics framework and accounts for the initial between analysis results for two laterally loaded piles with
sand fabric. It can closely capture peak strength, strain different diameters (B = 1 m and B = 10 m) using the two
softening, critical state and volumetric response (dilation and interface modelling approaches mentioned above. For the
contraction) of sand under different loading paths and initial contact pair approach, the pile surface was defined as the
stress states. Table 8 in Appendix 1 summarises the model master surface, and the soil surface was defined as the slave
equations and the values of the model parameters obtained surface. The ‘hard’ contact option was used in the normal

Table 1. Pile geometries, soil conditions and load eccentricities considered in the FE analyses

Pile diameter, Slenderness ratio, Diameter-to-wall thickness Sand type Relative Load eccentricity,
B: m L/B ratio, B/tw density, DR h: m

1–10 3–20 50–100 Ottawa sand, 40–80% 15–30


Toyoura sand

Table 2. Basic properties of Ottawa sand and Toyoura sand

Sand type Gs emax emin CU D50: mm ϕc,TXC: deg Particle shape

Ottawa sand 2·65 0·78 0·48 1·4 0·39 30·2 Rounded to subrounded
Toyoura sand 2·65 0·98 0·6 1·6 0·19 31·6 Angular

Note: Gs = specific gravity; emin = minimum void ratio; emax = maximum void ratio; CU = coefficient of uniformity; ϕc,TXC = critical-state
friction angle obtained from triaxial compression test.

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
LATERAL LOAD RESPONSE OF LARGE-DIAMETER MONOPILES IN SAND 1037

h = 15 m

L = 10 m
B=2m

10 m
Pile

Soil
40 m
m
20

Fig. 2. Mesh configuration for the 3D FE analysis of a monopile with B = 2 m, L = 10 m and h = 15 m

5000 used in this research given its many advantages, including


Mesh size
simpler model set-up, lower computational cost and more
stable convergence.
1 cm
4000 4 cm
10 cm
Lateral load at pile head: kN

Loading steps
Driving of large-diameter, open-ended pipe piles in sand
3000
takes place mostly in a fully coring mode (Han et al., 2020;
Lehane & Randolph, 2002; Paik et al., 2003), leading to only
small changes in the state of the sand surrounding the pile.
2000 Therefore, ‘wished-in-place’ piles under a fully coring
Ottawa sand
condition were assumed in the FE analyses.
B=2m
The FE analysis starts with an explicit step, during
L = 20 m which gravity – that is, weight ( = γ′V, where γ′ is the sand’s
1000
DR = 80% effective unit weight and V is the volume of elements being
h = 15 m considered) is applied to the pile–soil domain. In the same
step, initial vertical and horizontal effective stress fields
0 (σ′v0 = γ′z and σ′h0 = K0γ′z, where K0 is the coefficient of lateral
0 0·5 1·0 1·5 2·0 earth pressure) consistent with the gravity load are also
Pile rotation at mudline, θ : degrees applied to the domain. This step produces only negligibly
small displacement and deformation in the domain. Since
Fig. 3. Effect of mesh size on the load–rotation response of laterally
large-diameter monopiles are mostly used in offshore
loaded monopiles obtained from the FE analyses
conditions, sands are considered to be fully saturated in the
analyses reported in this paper. After the geostatic step, a
direction, whereas the Coulomb friction model was used in the horizontal velocity of 6 mm/s is applied at a height h above
tangential direction, with the interface shear stress τf calculated the mudline, as shown in Fig. 1. The velocity is smoothly
using τf = σn′ tanδcs, where σn′ is the normal effective stress and increased gradually from zero to 6 mm/s in the first 1 s, and
δcs is the pile–soil interface friction angle at critical state. then kept constant until the end of the loading.
According to Han et al. (2018) and Han et al. (2019a), the
interface friction angle δcs at critical state can be taken as 0·7ϕcs
for smooth steel and Ottawa sand with D50 = 0·39 mm. Given Validation of the analysis compared against the PISA load test
that steel piles in real applications are not smooth, this value In the PISA research project, lateral static load tests were
should maximise the difference between results obtained using performed on open-ended pipe piles with various diameters
the contact pair approach and the perfect contact approach. (0·273 m, 0·762 m and 2 m) installed in dense marine sand at
Because the critical-state friction angle of Ottawa sand is equal the Dunkirk test site (Byrne et al., 2019a). The test results of
to 30·2°, δcs = 21·14°. the largest test pile (B = 2 m) are used to validate the FE
As shown in Fig. 4, the results obtained using the two analyses in this paper. The test pile, which has a length
interface modelling approaches (in terms of the load– L = 10·5 m and a wall thickness tw = 38 mm (B/tw = 52), was
deflection response of the pile) are in close agreement, with loaded laterally at a height h = 10 m from the mudline. The test
the perfect contact approach producing slightly (2–3%) site consists of 3 m of sand fill placed in the 1970s, underlain by
greater lateral resistance for a pile rotation θ of 2° at the dense marine sand (Taborda et al., 2020; Zdravković et al.,
mudline. For the two piles with B = 1 m and B = 10 m, a pile 2020). According to Kuwano (1999), Dunkirk sand is a
rotation θ of 2° corresponds to pile deflection levels of 14%B uniform silica sand with rounded to subrounded particles with
and 9%B, respectively. The perfect contact approach was mean particle size D50 = 0·28 mm. The critical-state friction

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
1038 HU, HAN, PREZZI, SALGADO AND ZHAO
Pile deflection at mudline: mm Pile deflection at mudline: mm
0 20 40 60 80 100 120 140 160 0 200 400 600 800 1000
80 160 000

Interface modelled using perfect contact approach Interface modelled using perfect contact approach
Interface modelled using contact pair approach Interface modelled using contact pair approach

Lateral load at pile head: kN


120 000
Lateral load at pile head: kN

60

40 80 000

Ottawa sand Ottawa sand


B=1m B = 10 m
20 40 000
L=5m L = 50 m
h = 15 m h = 15 m
DR = 50% DR = 65%

0 0
0 0·5 1·0 1·5 2·0 2·5 0 0·5 1·0 1·5 2·0 2·5
Pile rotation at mudline, θ: degrees Pile rotation at mudline, θ: degrees
(a) (b)

Fig. 4. Load–deflection curves obtained by using the contact pair approach and the perfect contact approach for: (a) a 1 m dia., 5 m long pile in
sand with DR = 50% and (b) a 10 m dia., 50 m long pile in sand with DR = 65%

angle ϕc,TXC of Dunkirk sand is 32°. The groundwater table Fig. 5(c), the predicted and measured load–deflection curves
was found to be at a depth of 5·4 m at the time of the load test. at the mudline are in very close agreement.
Fig. 5(a) shows the cone penetration test (CPT) cone
resistances obtained at the test site (Zdravković et al., 2020).
The relative density was calculated using the equation
proposed by Salgado & Prezzi (2007) ANALYSIS RESULTS
Serviceability limit state criteria
lnðqc =PA Þ  04947  01041ϕcs  0841 lnðσ′h0 =pA Þ The design of wind turbine monopiles needs to pass
DR ð%Þ ¼ ultimate limit state (ULS) checks and serviceability limit
00264  00002ϕcs  00047 lnðσ′h0 =pA Þ
state (SLS) checks (Arany et al., 2017). Two key ULSs are
ð1Þ associated with (a) structural failure of the pile due to
exceedance of the yield moment at a pile cross-section or (b)
In initial calculations, a critical-state friction angle ϕcs = 32° geotechnical failure of the pile due to excessive pile rotation
(obtained from triaxial compression tests by Zdravković et al. and/or deflection. In contrast, SLSs are associated with
(2020)) and K0 = 0·4, corresponding to a normally consoli- relatively small pile rotations that are sufficiently large to
dated state, were used. The assumption that K0 = 0·4 led to compromise the functionality of the wind turbine even if it is
DR exceeding 100% from z = 0·5 m to 5·4 m, suggesting structurally sound. Table 4 summarises some of the ULS and
that the sand at the test site is in reality overconsolidated SLS criteria proposed in the literature. For the design of wind
within this depth range. The calculations were redone from turbine monopiles, SLS criteria are often critical, producing
z = 0·5 m to 5·4 m by fixing DR = 100% to obtain the K0 more conservative designs than those obtained using ULS
required to produce the measured qc values. The top 0·5 m criteria (Arany et al., 2015; LeBlanc et al., 2010).
of the ground may have been disturbed, unlocking some of Of the values listed in Table 4, a pile rotation θ at the
the lateral stresses, and thus the relative density calculations mudline equal to 0.5° has been commonly used as an SLS
with K0 = 0·4 produced DR = 76% for this layer. Table 3 criterion for offshore monopiles (Doherty & Gavin, 2012;
summarises the average DR and K0 estimated from all CPT Arany et al., 2017). However, when this value is mentioned in
data points (see Fig. 5(a)) measured in each soil layer. the literature, reference is usually made to the DNV offshore
The constitutive model used in the present study was standard (DNV GL, 2016), which indicated that a pile
calibrated by Loukidis & Salgado (2009) for Ottawa sand rotation angle at the mudline should be specified as an SLS
and Toyoura sand. Considering the similarities (in terms criterion by the offshore turbine manufacturer. The standard
of particle morphology and mineralogy) between Dunkirk provides θ = 0·5° only as an example. Given that the optimal
sand and Ottawa sand, the model parameters calibrated for value of θ to be used in routine design of these structures
Ottawa sand were used for Dunkirk sand, except for the value depends on the type of turbine, the piles are loaded to pile
of the slope Mcc of the critical state line (CSL) in q–p′ space, rotation values up to 2° in the FE simulations performed in
which is set to 1·28 for Dunkirk sand, corresponding to a this study.
critical-state friction angle = 32°. Using the updated model Since pile deflection is also used as an alternative to pile
parameters, isotropically consolidated, drained triaxial com- rotation to define limit states (as seen in Table 4), it is of
pression tests performed on Dunkirk sand with different interest for designers to know the relationship between pile
initial stress levels were simulated using the Loukidis & deflection and pile rotation at the mudline. Fig. 6 shows such
Salgado (2009) sand model. Fig. 5(b) shows a close match relationships for all the cases considered for Ottawa sand and
between the deviatoric stress plotted against axial strain Toyoura sand. As the example for a pile with B = 2 m,
curves obtained from the simulations and the experimental L = 20 m, h = 30 m and DR = 65% (represented by a thick
data reported in Taborda et al. (2020). The updated model solid line in Fig. 6(a)) shows, the normalised lateral pile
parameters were then used to simulate the PISA load test deflection u/(B 0·88L0·12
R ) at the mudline is nearly proportional
with the soil properties summarised in Table 3. As shown in to the pile rotation angle θ for values ranging from 0° to 1·5°,

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
LATERAL LOAD RESPONSE OF LARGE-DIAMETER MONOPILES IN SAND 1039
Cone resistance, qc: MPa
0 10 20 30 40 50 60
0 1600
Experimental data
2 1400 Simulation
Depth below the ground surface: m

p0' = 4
00 kP
1200 a, D
4 = 75%

Deviatoric stress, q: kPa


R

1000
6
800
8
600 p0' = 15
0 kPa,
DR = 7
10 Pile base 4%
400 p0' = 100 kPa,
DR = 73%
12 200 p0' = 50 kPa, DR = 73%

14 0
(a) 0 5 10 15 20
Axial strain, εa: %
(b)

6000
Load test result from PISA project
3D FE result
5000
Lateral load at pile head: kN

4000

3000

2000

B=2m
1000 h = 10 m
L = 10·5 m
tw/B = 1/52
0
0 50 100 150 200 250
Pile deflection at mudline: mm
(c)

Fig. 5. Validation of the FE analysis using the pile load test data from the PISA project: (a) the profile of the CPT cone resistance (data obtained
from Zdravković et al. (2020)); (b) comparison between experimental data (data obtained from Taborda et al. (2020)) and numerical simulation
results using the Loukidis & Salgado (2009) constitutive model for isotropically consolidated, drained triaxial compression tests performed on
Dunkirk sand; and (c) the load–deflection curves (data obtained from Taborda et al. (2020)) obtained from the static load test and the FE analysis

Table 3. Properties of the soil profile at the Dunkirk test site used in Table 4. Serviceability limit state (SLS) and ultimate limit state
the FE analysis (ULS) criteria defined in the literature for laterally loaded monopiles

Soil layer Depth: m Unit weight: K0 D R: % Reference SLS ULS


no. kN/m3
Luo et al. (2018) θ = 2°
1 0–0·5 17·1 0·4 76
Klinkvort & Hededal (2013) θ = 4°
2 0·5–3·0 17·1 2·5 100
Ahmed & Hawlader (2016) θ = 5°
3 3·0–4·6 17·1 1·3 100
Zdravković et al. (2015) u = 10%B
4 4·6–5·4 17·1 0·7 100
Doherty & Gavin (2012) θ = 0·5°
5 5·4–10·4 19·9 0·4 100
Arany et al. (2017) θ = 0·5° or u = 0·2 m
6 10·4– 19·9 0·4 92
Note: θ = pile rotation at the mudline; u = pile deflection at the mudline.
Note: The unit weight was obtained from Chow (1997) and
Zdravković et al. (2020). The water table was at 5·4 m below the
ground surface. Ottawa sand and Toyoura sand can be approximated as

R Þ ¼ 0064θ
u=ðB088 L012 ð2Þ
LR being the reference length ( = 1 m). In addition, as shown
in Fig. 6, all the cases considered fall in the range of According to equation (2), the pile deflections corres-
u/(B 0·88L0·12
R ) = 0·051θ and u/(B
0·88 0·12
LR ) = 0·077θ. Thus, the ponding to pile rotations equal to 0.5° and 1° at the mudline
average relationship between normalised pile deflection are equal to approximately 3·2%B 0·88 and 6·4%B 0·88,
R ) and pile rotation angle θ at the mudline for
u/(B 0·88L0·12 respectively.

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
1040 HU, HAN, PREZZI, SALGADO AND ZHAO
0·12 4000
All other cases considered in this study L = 40 m (L /I p0·25 = 68·20)
B = 2 m, L = 20 m, h = 30 m, DR = 65% L = 20 m (L /I p0·25 = 34·10)
Normalised pile deflection at mudline,
0·10
L = 15 m (L /I p0·25 = 25·58)
3000

Lateral load at pile head: kN


L = 10 m (L /I p0·25 = 17·05)
7 7θ
0·08 ·0
0 Ottawa sand
)=
u/(B0·88LR0·12)

12 B=2m

88 L
R h = 15 m
0·06 0· θ 2000 DR = 80%
(B 51
u/ ·0 tw/B = 1/50
2) =0
0·1 Ip = 0·118 m4
0·04 8LR
0·8
B
u/(
1000
0·02
Ottawa sand

0
0 0·5 1·0 1·5 0
0 40 80 120 160 200
Pile rotation at mudline, θ: degrees
(a) Pile deflection at mudline: mm
(a)
0·12
4000
All other cases considered in this study
L = 40 m (L /I p0·25 = 68·20)
B = 10 m, L = 40 m, h = 15 m, DR = 40%
L = 20 m (L /I p0·25 = 34·10)
Normalised pile deflection at mudline,

0·10
L = 15 m (L /I p0·25 = 25·58)
3000 L = 10 m (L /I p0·25 = 17·05)
Lateral load at pile head: kN
0·08 θ
0 77
u/(B0·88LR0·12)


12 )=
0·06 0·
88 LR
0· θ 2000
(B ·0 51
u/
0·04 2) =0
0·1
8LR
0·8
B
u/(
0·02 1000
Toyoura sand

0
0 0·5 1·0 1·5
Pile rotation at mudline, θ : degrees 0
(b) 0 0·5 1·0 1·5 2·0
Pile rotation at mudline: degrees
Fig. 6. Relationship between normalised pile deflection u/(B 0·88L0·12
R ) (b)
and pile rotation θ at the mudline for all cases considered for:
(a) Ottawa sand and (b) Toyoura sand Fig. 7. The effect of pile length L and pile length-to-area moment of
inertia ratio L/I0·25
p on (a) the load–deflection response and (b) the
load–rotation response for monopiles with B = 2 m (Ip = 0·118 m4) in
Critical pile length dense Ottawa sand (DR = 80%). The piles are loaded with an
It is expected that the lateral capacity of a pile with a fixed eccentricity h = 15 m
diameter increases with increasing embedded pile length.
Fig. 7 illustrates the evolution of the load–deflection response
of piles with B = 2 m installed in Ottawa sand with DR = 80% Table 5. Formulas available in the literature for estimation of the
and loaded with an eccentricity h = 15 m for increasing pile critical pile length
lengths. As L increases from 10 to 20 m, lateral resistance
mobilisation increases significantly, but further increases in Reference Method Equation
pile length (say from 20 to 40 m) result in negligible increases
in lateral resistance. This indicates the existence of a critical Davies & Elastic continuum kR = (EpIp)/EsL 4
pile length for lateral resistance mobilisation – that is, two Banerjee approach with the use
piles with the same diameter but with different lengths equal (1978) of integral equation
to or greater than the critical pile length behave in exactly solution
Poulos & Subgrade modulus Lcrit = 2·5
the same way when subjected to lateral loading (Davies Davis (1980) method (EpIp/khB)0·25
& Banerjee, 1978; Poulos & Davis, 1980; Randolph, 1981; Randolph Finite-element solution Lcrit = B(Eeq/G*)2/7
Fleming et al., 2008). A number of formulas have been (1981) for linear elastic soil
proposed in the literature that can be used to estimate
the critical pile length: they are summarised in Table 5. These Note: kR denotes the relative pile flexibility ratio (a pile is considered
infinitely long if kR is smaller than 104); EpIp is the pile bending
equations were obtained using the subgrade modulus
stiffness, where Ep is the Young’s modulus of pile and Ip is the area
methods (a simplified p–y method with linear elastic soil moment of inertia of the pile cross-section; Es is the Young’s
springs) or the FE method with a linear elastic soil model. modulus of soil; kh is the modulus of soil subgrade reaction; Eeq is
From Fig. 7(b), the lateral capacity H of the piles the equivalent Young’s modulus of the pile = EpIp/(πB 4/64); G* is the
corresponding to specific pile rotations (e.g. 0·5° or 1°) at modified shear modulus of soil = Gs(1 + 3/4vs), where Gs and vs are
the mudline can be determined. Fig. 8(a) plots the lateral the elastic shear modulus and Poisson’s ratio, respectively, of soil. G*
capacity H of the piles in Ottawa sand for θ = 0·5° and 1° as a increases linearly with increasing depth z: G* = mz.

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
LATERAL LOAD RESPONSE OF LARGE-DIAMETER MONOPILES IN SAND 1041
L/B Pile deflection: mm
0 5 10 15 20 25 –50 0 50 100 150
0
2400 SLS criterion: θ = 1°
SLS criterion: θ = 0·5°
Hcrit

Depth below the ground surface, z: m


2000
Lateral load capacity, H: kN

10

1600 L = 10 m (L /I p0·25 = 17·05)


L = 15 m (L /I p0·25 = 25·58)

Hcrit L = 20 m (L /I p0·25 = 34·10)


1200 20
L = 40 m (L /I p0·25 = 68·20)

800 Ottawa sand


B=2m B=2m
h = 15 m h = 15 m
400 DR = 80% 30 DR = 80%
tw/B = 1/50 tw/B = 1/50
Ip = 0·118 m4 Ip = 0·118 m4
0 SLS criterion: θ = 1°
0 10 20 30 40 50 60 70 80
L/I p0·25 40
–50 0 50 100 150
(a)

L/B Fig. 9. Effect of the pile length L or the L/I0·25


p ratio on the pile
deflection when θ = 1° for monopiles with B = 2 m (Ip = 0·118 m4) in
0 5 10 15 20 25
dense Ottawa sand (DR = 80%) loaded with an eccentricity h = 15 m
2400 SLS criterion: θ = 1°
SLS criterion: θ = 0·5°

2000
increases, the pile is more flexible and undergoes bending
Hcrit in addition to rotation. Pile bending continues to increase as
Lateral load capacity, H: kN

the pile length increases until it reaches a critical value of


1600 L = 20 m (L/I0·25p = 34·10 and L/B = 10), when the pile
deforms essentially like a clamped beam, with almost zero
1200
displacement at the pile base. When the pile length increases
Hcrit to 40 m (L/I0·25
p = 68·20 and L/B = 20), the deflection profile
of the pile for the top 20 m of the pile coincides with that of
800 Toyoura sand the 20 m long pile, and the deflection of the entire lower half
B=2m
h = 15 m
of the pile remains zero. The existence of the inactive pile
400 DR = 80% segment beyond the critical length explains the plateaus in
tw/B = 1/50 lateral capacity (the critical lateral resistance Hcrit) seen in
Ip = 0·118 m4 Fig. 8. For monopile design considerations, increasing the
0 pile length beyond the critical length is not a cost-effective
0 10 20 30 40 50 60 70 80
option as far as lateral pile capacity is concerned.
L/I p0·25
Figure 10 compares the soil displacement for 2 m dia.
(b) monopiles with two different lengths: L = 10 m (L/B = 5) and
L = 40 m (L/B = 20). Both piles are embedded in dense
Fig. 8. Effect of L/B and L/I0·25 p on the lateral capacity H Ottawa sand (DR = 80%) and loaded at h = 15 m to a pile
corresponding to θ = 0·5° and 1° for monopiles with B = 2 m
rotation of 1° at the mudline. As the short pile undergoes
(Ip = 0·118 m4) in dense sand (DR = 80%) loaded with an eccentricity
h = 15 m for: (a) Ottawa sand and (b) Toyoura sand rigid-body rotation (the case with L/B = 5), the sand mass
(inside and around the pile) rotates about a rotation centre
located at a depth of approximately 7 m. Behind the short
pile, owing to the rigid-body rotation, the sand above the
function of the slenderness ratio L/B or the non-dimensional rotation centre undergoes unloading (reflected as a decrease
pile length-to-area moment of inertia ratio L/I0·25 p . in the mean stress, as shown in Fig. 11(a)), whereas
Ip = π(B 4B4i )/64 is the area moment of inertia of the pile significant pressure build-up is seen in the sand below the
cross-section, and Bi is the inner diameter of the pipe pile. As rotation centre. In contrast, in front of the pile, loading and
shown in Fig. 8(a), the lateral capacity H (corresponding to unloading are observed in the sand above and below the
θ = 0·5° or 1°) for piles with a fixed diameter increases with rotation centre, respectively. In addition, as the pile toe kicks
increasing pile length until L reaches a critical value Lcrit, back, it pushes on the sand plug (the sand volume inside
when it stabilises at a critical lateral capacity Hcrit. Similar the pile), leading to an increase of the mean stress there
trends regarding the critical pile length are also observed (Fig. 11(a)) and to significant shearing between the sand plug
for Toyoura sand, as shown in Fig. 8(b). and the sand below the pile base (as shown in Fig. 11(b)).
In order to understand the effect of the pile length on For the long pile case (the case with L/B = 20), in contrast,
the pile response to lateral loading and the reason for the sand generally undergoes loading ahead of the pile and
the existence of a critical lateral capacity Hcrit, the lateral unloading behind it (Fig. 11(c)), while minimal interaction
deflection profiles of the piles in Ottawa sand considered is observed between the pile and the sand plug. The critical
in Figs 7 and 8 are plotted in Fig. 9 for θ = 1°. When the pile length can also be identified from the stress field in the
pile slenderness ratio is small (L/B = 5 and L/I0·25
p = 17·05), as sand around the pile (see Fig. 11(c)); the mean stress in the
is the case for L = 10 m, the pile rotates as a rigid body. As the sand is affected only by the pile segment above about 20 m in
pile length increases, and hence, the slenderness ratio depth ( = 10B), below which a geostatic soil state exists

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
1042 HU, HAN, PREZZI, SALGADO AND ZHAO

Displacement 0 0
in soil: m
0·128 2 2

Depth below the mudline: m


Depth below the mudline: m
0·117
0·107 4 4
0·096
0·085
0·075 6 6
0·064
0·053 8 8
0·043 B=2m B=2m
0·032 L = 10 m 10 L = 40 m
10
0·021
h = 15 m h = 15 m
0·011
0 12 DR = 80% 12 DR = 80%

(a) (b)

Fig. 10. Contour and vector plots of the soil displacement for 2 m dia. monopiles with: (a) L = 10 m (L/B = 5); and (b) L = 40 m (L/B = 20), in
dense Ottawa sand (DR = 80%) loaded at h = 15 m to a pile rotation of 1° at the mudline. The pile deflection is magnified by 5 times in the figure
for illustration purposes
Depth below the mudline, z: m

0 0

Mean stress:
5 5 kPa
400
10 Ottawa sand 10 367

Depth below the mudline, z: m


333
D=2m 300
15 L = 10 m 15
267
233
h = 15 m 200
DR= 80% 167
20 20 133
(a) 100
67
33
25 0
Plastic dissipation
Dp: kJ/(m3 s)
30
30
26 Ottawa sand
23 35
19 D=2m
15 L = 40 m
11 h = 15 m 40
8 DR= 80%
4
0
(b) (c)

Fig. 11. Contour plots of (a) mean stress; (b) plastic dissipation in the sand for a 2 m dia. monopile with L = 10 m (L/B = 5); (c) mean stress in
sand for a 2 m dia. monopile with L = 40 m (L/B = 20). Both piles are embedded in dense Ottawa sand (DR = 80%) and loaded at h = 15 m to a
pile rotation of 1° at the mudline. The pile deflection is magnified by 5 times in the figure for illustration purposes

(observed as horizontal contour lines), indicating no pile–soil the H/Hcrit ratio approaches 1·0 at a greater rate with
interaction. respect to L/B, leading to a smaller critical slenderness
The local interaction between the pile and the sand around ratio Lcrit/B. This is because the stiffness of the sand increases
it and inside it, as well as the interaction between the sand with depth due to increasing confining stress. Thus, the
plug and the sand below it – all combined – control the greater the depth is, the more difficult it is for the pile to
response of monopiles to lateral loading. In this sense, 3D FE deflect laterally. The zero-deflection point is reached further
analyses using a realistic constitutive model are able to up the pile for a pile with larger diameter. Note that
capture the complex pile–soil interaction processes (as shown the Lcrit/B ratio decreases with increasing pile diameter B;
in Figs 10 and 11) and produce results that are more complete however, the critical pile length Lcrit itself increases with
and of higher quality than the p–y method, which does not increasing pile diameter B.
treat the soil surrounding the pile as a continuum and, in In addition to diameter, the sand relative density DR also
particular, treats the pile as a beam with no consideration of plays an important role in the relationship between the lateral
the soil plug. capacity ratio H/Hcrit and the L/I0·25
p ratio (or the slenderness
It is convenient to work with the lateral capacity ratio ratio L/B). Fig. 13 shows the relationship between H/Hcrit
H/Hcrit. Figs 12(a) and 12(b) show the relationships between and L/I0·25
p for a 2 m dia. pile in Ottawa sand and Toyoura
the lateral capacity ratio H/Hcrit and the L/I0·25
p ratio (or the sand with three different relative densities. The lateral
L/B ratio) in dense (DR = 80%) Ottawa sand and Toyoura capacity H of a 2 m dia. pile approaches its critical value
sand, respectively, for piles with three different diameters Hcrit at a greater rate with respect to L/B as DR increases.
(1 m, 4 m and 10 m). The trends of H/Hcrit against L/I0·25 p Therefore, the critical length Lcrit decreases with increasing
for the two sands are almost identical. The lateral capacity relative density, all else being equal. The reason for this is the
ratio H/Hcrit increases as the normalised pile length – in gain in stiffness from the increasing relative density; this
terms of L/B or L/I0·25
p – increases, stabilising at 1·0 when L/B enables the pile to absorb the load and moment at its top by
becomes greater than the critical value Lcrit/B. With increas- mobilising forces along a shorter length. In contrast, as
ing pile diameter B (and, consequently, increasing Ip), shown in Fig. 13, the load eccentricity h and sand type have

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
LATERAL LOAD RESPONSE OF LARGE-DIAMETER MONOPILES IN SAND 1043
a negligible effect on the relationship between the lateral relative density DR and the area moment of inertia Ip of the
capacity ratio H/Hcrit and the L/I0·25
p ratio (or the L/B ratio). pile’s cross-section (as demonstrated in Figs 12 and 13)
Based on all the data obtained from the FE analyses of the !
 Ip025
cases summarised in Table 1, the fitted relationship between 025  DR
Lcrit =Ip  ¼ 103  35 ln þ 37
the lateral capacity ratio H/Hcrit and the L/I0·25
p ratio for both 05° 100% LR ð4Þ
Ottawa sand and Toyoura sand can be expressed as 
8 
βðDR ; Ip Þ 05° ¼ 0084ðLcrit =Ip Þ  12
025
  β
>
< 1 sin π L  π þ 05 ; when L , L
crit for mudline pile rotation θ = 0·5°, and
H=Hcrit ¼ 2 Lcrit 2
>
: !

Ip025
1; when L  Lcrit  DR
Lcrit =Ip025  ¼ 103  32 ln þ 394
ð3Þ 1° 100% LR ð5Þ
  
where Lcrit is the critical length of the pile. According to βðDR ; Ip Þ1° ¼ 008 Lcrit =Ip025  12
equation (3), the value of the H/Hcrit ratio increases from 0 to
1 as the pile length increases from 0 to the critical pile length for mudline pile rotation θ = 1°. LR is the reference
Lcrit, and then stabilises at 1 with further increase in pile length = 1 m = 3·281 ft.
length. The parameter β in equation (3) determines the rate at
which the H/Hcrit ratio approaches 1; it is a function of the
L /B
L /B 0 5 10 15
0 5 10 15
1·0
1·0

Lateral capacity ratio, H/Hcrit


0·8
Lateral capacity ratio, H/Hcrit

0·8
DR = 80%, h = 15 m
DR = 80%, h = 30 m
0·6
0·6 DR = 65%, h = 15 m
B = 10 m, Ip = 73·95 m4
DR = 65%, h = 30 m
B = 4 m, Ip = 1·893 m4
DR = 50%, h = 15 m
0·4
0·4 B = 1 m, Ip = 0·007 m4
DR = 50%, h = 30 m
Obtained using equation (3) Obtained using equation (3)
0·2
0·2 Ottawa sand
Ottawa sand
B = 2 m, Ip = 0·118 m4
h = 15 m, DR = 80%
tw /B = 1/50, pile rotation, θ = 0·5°
tw /B = 1/50, pile rotation, θ = 0·5°
0 0
0 10 20 30 40 50 0 10 20 30 40 50

L/I p0·25 L/Ip0·25

(a) (a)

L /B
L /B
0 5 10 15
0 5 10 15

1·0
1·0
Lateral capacity ratio, H/Hcrit

0·8
Lateral capacity ratio, H/Hcrit

0·8
DR = 80%, h = 15 m
DR = 80%, h = 30 m
0·6 0·6
B = 10 m, Ip = 73·95 m4 DR = 65%, h = 15 m
B = 4 m, Ip = 1·893 m4 DR = 65%, h = 30 m

B = 1 m, Ip = 0·007 m4 0·4 DR = 50%, h = 15 m


0·4
Obtained using equation (3) DR = 50%, h = 30 m
Obtained using equation (3)
0·2 0·2
Toyoura sand Toyoura sand
h = 15 m, DR = 80% B = 2 m, Ip = 0·118 m4
tw /B = 1/50, pile rotation, θ = 0·5° tw /B = 1/50, pile rotation, θ = 0·5°
0 0
0 10 20 30 40 50 0 10 20 30 40 50
L/I p0·25 L/Ip0·25
(b) (b)

Fig. 12. Effect of pile diameter B on the relationship between the Fig. 13. Effect of the relative density DR and the load eccentricity h
lateral capacity ratio H/Hcrit and the L/I0·25
p ratio or the slenderness on the relationship between the lateral capacity ratio H/Hcrit and the
ratio L/B for monopiles in: (a) dense (DR = 80%) Ottawa sand and L/I0·25
p ratio for pile with B = 2 m obtained for θ = 0·5° at the mudline
(b) dense Toyoura sand, loaded with an eccentricity h = 15 m for for: (a) Ottawa sand and (b) Toyoura sand. Dotted curves are obtained
θ = 0·5°. Dotted curves are obtained from equation (3) from equation (3)

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
1044 HU, HAN, PREZZI, SALGADO AND ZHAO
Table 6. Critical slenderness ratio Lcrit/B calculated using the proposed method

Case B=1 m B=1 m B=4 m B=4 m B = 10 m B = 10 m


DR = 50% DR = 80% DR = 50% DR = 80% DR = 50% DR = 80%

Proposed method at θ = 0·5° 10·6 9·7 9·2 8·3 7·1 6·4


Proposed method at θ = 1° 11·2 10·3 9·9 9·0 7·8 7·0

Note: Wall thickness tw/B = 1:50 for piles of B = 1 m and B = 4 m; wall thickness tw/B = 1:100 for piles of B = 10 m.

Figures 12 and 13 show the quality of the fit of the 16 000


equations (shown as dotted lines in the figures) to the data h = 15 m
points obtained from the FE analyses. As an illustration of h = 30 m
the values produced by the proposed method, Lcrit values

Critical lateral load capacity, Hcrit: kN


Ottawa sand
were calculated for a few representative cases. As shown in 12 000 B=4m
Table 6, the critical slenderness ratio Lcrit/B for θ = 0·5° tw /B = 1/50
decreases from 10·6 for a 1 m dia. pile to 6·4 for a 10 m dia. Ip = 1·893 m4
pile. Pile rotation at
mudline: θ = 0·5°
The equations in Table 5 were derived based on linear 8000
elasticity. They require the estimation of either a representa-
tive elastic modulus for the soil around the pile or the rate at
which the elastic modulus increases with depth. These are
difficult to estimate for two reasons: (a) soil is not elastic and 4000
(b) if an elastic modulus is used to represent the stress–strain
response of soil, that modulus must change with soil
deformation. Thus, using these equations to estimate Lcrit
accurately in practice is challenging. In contrast, with the 0
30 40 50 60 70 80 90
proposed methodology, such estimates are not required. The
Relative density: %
proposed equation for Lcrit is a function of the relative (a)
density DR and the area moment of inertia Ip of the pile’s
cross-section. 28 000
h = 15 m
24 000 h = 30 m
Critical lateral load capacity, Hcrit: kN

Ottawa sand
B=4m
Critical lateral load capacity 20 000 tw /B = 1/50
The lateral capacity H of a pile with a certain cross-section Ip = 1·893 m4
increases with increasing pile length L and reaches the 16 000 Pile rotation at
critical lateral capacity Hcrit when L becomes equal to the mudline: θ = 1°
critical length Lcrit. The critical lateral capacity Hcrit is a
12 000
function of: the pile cross-section (i.e. Ip); sand type and
relative density DR; load eccentricity h; and the rotation level
that is used to obtain the capacity. Figs 14(a) and 14(b) 8000
show the effect of relative density and load eccentricity on
Hcrit of 4 m dia. piles (Ip = 1·893 m4) for θ = 0·5° and 4000
θ = 1·0°. The critical lateral capacity Hcrit increases linearly
with increasing relative density, regardless of the load 0
eccentricity and the value of θ. The greater the load 30 40 50 60 70 80 90
eccentricity h is, the greater is the bending moment applied Relative density: %
on the pile; higher h values lead to smaller critical lateral (b)
capacity Hcrit.
The effects of pile diameter B (or Ip) on the critical lateral Fig. 14. Effect of relative density DR and load eccentricity h on the
pile critical lateral load capacity Hcrit for piles with B = 4 m placed in
capacity Hcrit are illustrated in Fig. 15 for the case of Ottawa sand. The value of Hcrit corresponds to (a) mudline pile
medium-dense (DR = 50%) Ottawa sand. The critical lateral rotation θ = 0·5° and (b) mudline pile rotation θ = 1·0°
capacity Hcrit increases almost cubically with increasing pile
diameter B (or I0·25
p ). The significant increase of Hcrit is not
only attributed to the direct impact of increasing the pile increasing B, as the contact area (πBL) between the pile
diameter B (and the area of the pile in contact with soil), and soil increases linearly with B.
but also to the correspondingly greater critical pile length Based on the relationships observed in Figs 14 and 15 and
Lcrit required for Hcrit to be mobilised. For example, Hcrit the dataset obtained from the FE analyses, an equation
is achieved at Lcrit equal to about 40 m for a pile with was derived to estimate the critical lateral capacity Hcrit as
B = 4 m, whereas Lcrit is only about 11 m for a pile with follows
B = 1 m. For comparison, the load–rotation curves for three  072
piles with the same length (L = 10 m) but different diameters Hcrit Ip
(B = 1 m, 2 m and 4 m) in Ottawa sand are plotted in ¼ aðDR ; hÞ þb ð6Þ
pA L2R L4R
Fig. 16(a). The lateral capacities H of these piles correspond-
ing to θ = 1° obtained from Fig. 16(a) are plotted as a The form of equation (6) captures the approximately
function of B in Fig. 16(b). For both relative densities cubical increase of the critical lateral capacity Hcrit with
(DR = 50% and 80%), H increases nearly linearly with increasing pile diameter B (or I0·25 0·72
p ). Note that the term Ip

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
LATERAL LOAD RESPONSE OF LARGE-DIAMETER MONOPILES IN SAND 1045
Pile diameter, B: m 2000
0 4 8 12
100 000 B=4m
h = 15 m B=2m
Critical lateral load capacity, Hcrit: kN

h = 30 m B=1m

Lateral load at pile head: kN


80 000
L = 10 m
Ottawa sand h = 30 m
60 000 DR = 50% DR = 80%
1084 kN
tw /B = 1/50 1000 tw /B = 1/50
Pile rotation at
40 000 mudline: θ = 0·5°

510 kN

20 000
183 kN

0
0 1 2 3 0
0 0·5 1·0 1·5 2·0
Area moment of inertia of the cross-section, I 0·25
p :m
Pile rotation at mudline: degrees
(a)
(a)
Pile diameter, B: m
0 4 8 12 2000
16 000
h = 15 m DR = 80%
h = 30 m DR = 50%
Critical lateral load capacity, Hcrit: kN

12 000
Lateral load capacity, H: kN
Ottawa sand
DR = 50% L = 10 m
tw /B = 1/50 h = 30 m
Pile rotation at tw /B = 1/50
8000
mudline: θ = 1° 1000 Pile rotation at
mudline: θ = 1°

4000

0
0 1 2 3
Area moment of inertia of the cross-section, I 0·25
p :m 0
0 1 2 3 4 5
(b)
Pile diameter, B: m
Fig. 15. Critical lateral load capacity Hcrit of piles embedded in (b)
medium dense sand (DR = 50%) with different pile diameters and load
eccentricities. The value of Hcrit is determined for two values of pile Fig. 16. Effect of the pile diameter B on the lateral capacity H for
rotation at the mudline: (a) pile rotation θ = 0·5° and (b) pile rotation piles with the same length: (a) load–rotation curves for three piles with
θ = 1·0° L = 10 m in dense Ottawa sand (DR = 80%); (b) H obtained at θ = 1°
plotted against pile diameter for piles with L = 10 m in medium dense
(DR = 50%) and dense (DR = 80%) Ottawa sand

in the equation is close to (I0·25 3 0·75


p ) = Ip . The term a(DR, h)
captures the dependence of Hcrit on the relative density DR Table 7. Values of the coefficients in equations (6) and (7) to calculate
and load eccentricity h the critical lateral capacity Hcrit in Ottawa sand and Toyoura sand for
  θ = 0·5° and θ = 1°
h DR h
aðDR ; hÞ ¼ a1 þ a2 þ a3 þ a4 ð7Þ Case a1 a2 a3 a4 b
LR 100% LR
In equations (6) and (7), the coefficients a1–a4 and b are Ottawa sand for 0·84 46·69 0·34 25·64 0·02
obtained by fitting the results of the 3D FE analyses for θ = 0·5°
Ottawa sand and Toyoura sand for θ = 0·5° and θ = 1°. Ottawa sand for 1·86 102·36 0·31 32·35 0·07
θ = 1°
Table 7 summarises the values of these coefficients.
Toyoura sand for 0·61 34·94 0·48 31·72 0·03
The lateral load capacity H of a monopile can be estimated θ = 0·5°
by calculating the H/Hcrit ratio using equation (3) and the Toyoura sand for 1·25 70·87 0·62 46·28 0·09
value of Hcrit using equations (6) and (7). To verify the θ = 1°
accuracy of the proposed equations, the lateral capacities
for all the cases considered in the present study (for both
Ottawa sand and Toyoura sand) were calculated using the
proposed equations (equations (3)–(7)) and compared with thickness-to-diameter ratios tw/B ranging from 1:100 to
those obtained from the 3D FE analyses in Fig. 17. Excellent 1:50, two types of sands (Toyoura sand and Ottawa sand)
agreement is observed for these predictions. The proposed with different particle morphologies, relative densities
method is applicable to pile diameters ranging from 1 to ranging from 40 to 80%, and load eccentricities ranging
10 m, slenderness ratios L/B ranging from 3 to 20, wall from 15 to 30 m.

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
1046 HU, HAN, PREZZI, SALGADO AND ZHAO
1 000 000 80

fit
ct
B = 10 m

fe
5%

r
3D FE analysis

Pe
B=8m

+1
Lateral load capacity estimated

100 000
B=4m Proposed equation
from proposed method: kN

5%
B=2m 60

–1

Lateral load at pile head: kN


B=1m
10 000

1000 40

Ottawa sand
100 Ottawa sand B=1m
pile rotation at L=5m
mudline: θ = 1° 20 h = 30 m
DR = 80%
10
10 100 1000 10 000 100 000 1 000 000 tw/B = 1:50

Lateral load capacity obtained from


FE analyses: kN 0
(a) 0 0·5 1·0 1·5
Pile rotation at mudline: degrees
1 000 000 fit (a)
ct

B = 10 m
fe
5%

80 000
Pe

B=4m
+1
Lateral load capacity estimated

100 000
B=2m
from proposed method: kN

5%

3D FE analysis
B=1m
–1

Proposed equation
10 000 60 000
Lateral load at pile head: kN

1000

40 000
100 Toyoura sand
pile rotation at Ottawa sand
mudline: θ = 1° B = 10 m
L = 80 m
10 20 000 h = 30 m
10 100 1000 10 000 100 000 1 000 000 DR = 40%
Lateral load capacity obtained from tw/B = 1:100
FE analyses: kN
(b)
0
0 0·5 1·0 1·5
Fig. 17. Comparison of lateral capacities H estimated using the
Pile rotation at mudline: degrees
proposed equations with those obtained from the FE analyses for all
(b)
cases considered in the present study corresponding to pile rotation
θ = 1° for: (a) Ottawa sand and (b) Toyoura sand
Fig. 18. Lateral load–rotation responses obtained from FE analyses
and using equations (8) and (9) for: (a) a 1 m dia. pile with L = 5 m,
h = 30 m in dense (DR = 80%) Ottawa sand; and (b) a 10 m dia. pile
Prediction of the lateral load–rotation curve with L = 80 m, h = 30 m in loose (DR = 40%) Ottawa sand
In addition to the estimation of the lateral capacity of a
monopile at specific rotation values (θ = 0·5° and 1°), it is useful
to have the lateral load–rotation curve as well. According to Excellent agreement is found between the two methods.
Fig. 7, the lateral pile capacity H increases with increasing For convenient application of the proposed method
rotation θ at the mudline following a power function to calculate the lateral pile capacity and to obtain the
load–rotation curve, a web-based application was developed
θ
H¼ ð8Þ based on the results of the FE simulations. The link to the
k þ ηθ web-based application is provided in Appendix 2.
where the coefficients k and η control the magnitude of H
and the rate at which H increases with θ. Since there are two
unknown coefficients, only two points on the curve (H Comparison of the lateral load capacities of piles in Ottawa
plotted against θ) are needed to determine the values of k and sand and Toyoura sand
η. This can be achieved by calculating Hθ=0·5° and Hθ=1° for In Fig. 19, the lateral capacity HToyoura of a 4 m dia. pile in
θ = 0·5° and θ = 1° using equations (3)–(7). Substituting Toyoura sand is compared with the lateral capacity HOttawa of
(Hθ=0·5°, θ = 0·5°) and (Hθ=1°, θ = 1°) into equation (8), the the same pile in Ottawa sand. This comparison is done for
following is obtained piles with different lengths and sands with different relative
densities. The lateral capacity HToyoura is slightly greater than
η ¼ 2=Hj1°  1=Hj05° or equal to HOttawa when the relative density DR is small (e.g.
ð9Þ
k ¼ 1=Hj05°  1=Hj1° DR = 40%). The ratio of HToyoura to HOttawa decreases as the
relative density DR increases. The differences in the lateral
Figure 18 compares the load–rotation curves obtained capacities in these two sands may be attributed to the fact
from the 3D FE analyses and those calculated using that they have different particle morphology and initial
equations (8) and (9) for two different piles: a short fabric, which affects dilatancy. The HToyoura/HOttawa ratio is
small-diameter pile and a long large-diameter pile. independent of the pile length. This is consistent with the

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
LATERAL LOAD RESPONSE OF LARGE-DIAMETER MONOPILES IN SAND 1047
L /B the load eccentricity h increases, the bending moment also
0 4 8 12 increases, leading to smaller critical lateral pile capacity Hcrit.
1·2
Equations were proposed to calculate the lateral capacities
DR = 40% Hθ=0·5° and Hθ=1° of a monopile for pile rotations θ = 0° and
DR = 50% θ = 1° at the mudline. In addition, equations were also
DR = 65%
proposed to obtain the load–rotation curve for a pile, so
1·1 that its lateral capacity can be obtained for a rotation level
DR = 80% ranging from 0° to 2°, as is needed in design. The proposed
equations are applicable to a broad range of pile diameters B
HToyoura/HOttawa

from 1 to 10 m, slenderness ratios L/B from 3 to 20, wall


1·0
thickness-to-diameter ratios tw/B from 1:100 to 1:50, two
types of sands (Toyoura sand and Ottawa sand) with different
particle morphologies, relative densities from 40 to 80% and
load eccentricities from 15 to 30 m. Readers can choose from
the proposed equations by comparing the properties of the
0·9 sand of interest to those (Toyoura sand and Ottawa sand)
B=4m
h = 15 m considered in this study.
tw /B = 1:50 The equations proposed in this study are applicable to: (a)
θ = 0·5° monotonic lateral loading conditions; (b) uniform sand
0·8 profiles; and (c) input parameters in the range of those
0 10 20 30 40 considered in the analyses. The effect of cyclic loading on the
L/I 0·25
p static lateral capacity of monopiles has not been considered
in the development of the proposed equations.
Fig. 19. Ratio of the lateral load capacity in Toyoura sand to that in
Ottawa sand for a 4 m dia. pile with tw/B = 1:50. The lateral load
capacities are obtained for θ = 0·5°
APPENDIX 1
See Table 8.
earlier observation that the relationship between the H/Hcrit
ratio and the slenderness ratio L/B is independent of the sand
type. APPENDIX 2
The web application can be found through the link provided in the
references (Purdue University, 2021).

CONCLUSIONS NOTATION
A series of 3D FE analyses was performed to investigate B outer diameter of pile
the response of monopiles subjected to lateral loads using an Bi inner diameter of pile
advanced two-surface-plasticity constitutive model. Two DR relative density of sand critical-state friction angle
types of sand (Ottawa sand and Toyoura sand) and a wide Ep Young’s modulus of pile
range of pile diameters, slenderness ratios and wall thick- h load eccentricity (the distance from the lateral load to
nesses, load eccentricities and sand relative densities were mudline)
considered in the FE analyses. H lateral capacity of pile
The lateral capacity H of a pile, for a fixed diameter, Hcrit critical lateral capacity of pile
Ip area moment of inertia of pile
increases with increasing pile length, but only while the K0 coefficient of the lateral earth pressure
length does not exceed a critical pile length Lcrit, beyond L pile length
which H stops increasing and stabilises at a critical value Lcrit critical pile length
Hcrit. The critical slenderness ratio Lcrit/B decreases as the LR reference length = 1 m
pile diameter B or the sand relative density DR increases, but pA reference stress = 100 kPa
this ratio has a negligible dependence on the sand type. The qc cone penetration test cone resistance
critical slenderness ratio Lcrit/B decreases from about 11 to 6 tw wall thickness of pile
when the pile diameter increases from 1 m to 10 m. An u pile deflection at mudline
equation was proposed that can be used to estimate the δcs critical-state interface friction angle
θ pile rotation at mudline
critical pile length Lcrit for both Ottawa sand and Toyoura
ν Poisson’s ratio of pile
sand. Accurate estimation of Lcrit is important because a σ′h0 in situ horizontal effective stress in soil = K0 σ′v0
design for a laterally loaded pile with L greater than Lcrit is σ′n normal effective stress at pile–soil interface
not economical. The critical lateral pile capacity Hcrit σ′v0 in situ vertical effective stress in soil
increases almost cubically with increasing pile diameter τf interface shear stress = σ′n tanδcs
and increases linearly with increasing relative density. As ϕcs critical-state friction angle

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
1048
Table 8. Constitutive model equations and values of the model parameters for Ottawa sand and Toyoura sand

Description Constitutive equations Parameters Parameter value

Ottawa Toyoura
sand sand
 
2
Stress–strain relations σ̇ij′ ¼ 2Gðε̇ij  ε̇pij Þ þ K  G ðε̇kk  ε̇pkk Þ — — —
3
ð217  eÞ2 ′ 1ng
Small-strain shear modulus Gmax ¼ Cg p ng pA Cg 611 900
1þe ng 0·437 0·400
Gmax Gmax
Elastic shear modulus with Ramberg–Osgood G¼ n pffiffiffiffiffiffiffiffipffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi o α1 0·47 0·40
1 þ 2½ð1=α1 Þ  1 ½ 3=2 ðrij  aij;ini Þðrij  aij;ini Þ=½2α1 ðGmax = p′ Þγ1  1 þ 2½ð1=α1 Þ  1
degradation γ1 0·00065 0·0010
2ð1 þ νÞ
K¼ ν

HU, HAN, PREZZI, SALGADO AND ZHAO


Elastic bulk modulus G 0·15 0·15
3ð1  2νÞ
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi
Yield surface f ¼ ðsij  aij p′ Þðsij  aij p′ Þ  2=3m p′ ¼ 0 m 0·05 0·05
 
1 1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Plastic multiplier Λ̇ ¼ nij  ðnkl rkl Þ  δij σ̇ij′ where nij ¼ ðrij  aij Þ= ðrij  aij Þðrij  aij Þ and rij ¼ sij = p′ — — —
H 3
Bounding, dilatancy and CS surfaces Mb ¼ gðθÞ  Mcc exp½kb ψ , Md ¼ gðθÞ  Mcc exp½kd ψ , Mc ¼ gðθÞ  Mcc Mcc 1·21 1·27
kb 1·9 1·5
kd 2·2 2·8
ξ
State parameter and CSL in e–p′ ψ ¼ e  ec , ec ¼ Γ  λ p′ =pA λ 0·081 0·019
ξ 0·20 0·70
1=n 1=n
Shape of bounding, dilatancy, CS surfaces in the π gðθÞ ¼ 2ns c1 =½1 þ c1 s  ð1  c1 s Þ cosð3θÞns c1 0·71 0·72
plane rffiffiffi pffiffiffiffiffiffiffiffi   ns 0·35 0·35
2 G  exp½kb ψ   ð 2=3ðMb  mÞ  aij nij Þ elim  e h1
Plastic modulus H ¼ h 0 hk  hpffiffiffiffiffiffiffiffiqffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiiμ , h0 ¼ h1 2·20 1·62
3 3=2 r  a r a h2 h2 0·240 0·254
ij ij;ini ij ij;ini
elim 0·81 1·0
  μ 1·2 2·0
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Flow rule ε̇pij ¼ Λ̇  Rij′ þ Dδij , R′ij ¼ Rij′= Rkl′Rkl′ c2 0·78 0·78
3
    " rffiffiffi  #  
 3 1  c2 3 1  c2 1
Rij ′ 1 þ g2 ðθÞ cosð3θÞ nij  3 g2 ðθÞ  nik nkj  δij
2 c2 2 c2 3
rffiffiffi !
D0 2
Dilatancy D¼  ðMd  mÞ  aij nij D0 1·31 0·90
Mcc 3
Shape of plastic potential in π plane g2 ðθÞ ¼ 2c2 =½ð1 þ c2 Þ  ð1  c2 Þ cosð3θÞ — — —
Fabric tensor F 11 ¼ F 33 ¼ 05ð1  αÞ; F 22 ¼ α; F ij ¼ 0 for i = j α 0·31 0·29
Fabric-dependent scalar Af ¼ gðθÞ F ij nij  pffiffiffiffiffiffiffiffi
Fabric effect multiplier in H hk ¼ expf½ðAfc  Af Þ=ðAfc  Afe Þ125  lnðkh Þg, Afc ¼ ð1=c1 ÞAfe ¼ 3=2ðα  1=3Þ kh 0·39 0·11
Intercept of CSL Γ ¼ Γc  expðAfc  Af Þ Γc 0·780 0·934

Note: G = current value of shear modulus; K = bulk modulus; H = plastic modulus; σ′ij = effective stress tensor; sij = deviatoric stress tensor; p′ = mean effective stress; aij = kinematic hardening tensor;
aij,ini = initial value of kinematic hardening tensor; δij = Kronecker’s delta tensor; εij = strain tensor; Λ̇ = plastic multiplier; εijp = plastic strain tensor.

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
LATERAL LOAD RESPONSE OF LARGE-DIAMETER MONOPILES IN SAND 1049
REFERENCES Han, F., Ganju, E., Salgado, R. & Prezzi, M. (2018). Effects of
Abaqus (2014). Abaqus 6.14-4. Abaqus analysis user’s manual. interface roughness, particle geometry, and gradation on the
Providence, RI, USA: Simulia Inc. sand–steel interface friction angle. J. Geotech. Geoenviron.
Ahmed, S. S. & Hawlader, B. (2016). Numerical analysis of Engng 144, No. 12, 04018096.
large-diameter monopiles in dense sand supporting offshore Han, F., Ganju, E., Prezzi, M. & Salgado, R. (2019a).
wind turbines. Int. J. Geomech. 16, No. 5, 04016018. Closure to ‘Effects of interface roughness, particle
API (American Petroleum Institute) (2014). Recommended geometry, and gradation on the sand–steel interface friction
practice for planning, designing and constructing fixed offshore angle’ by Fei Han, Eshan Ganju, Rodrigo Salgado, and Monica
platforms load and resistance factor design. Washington, DC, Prezzi. J. Geotech. Geoenviron. Engng 145, No. 11, 07019017.
USA: API. Han, F., Salgado, R., Prezzi, M. & Lim, J. (2019b). Axial resistance
Arany, L., Bhattacharya, S., MacDonald, J. H. G. & Hogan, S. J. of nondisplacement pile groups in sand. J. Geotech. Geoenviron.
(2015). A critical review of serviceability limit state requirements Engng 145, No. 7, 04019027.
for monopile foundations of offshore wind turbines. Proceedings Han, F., Ganju, E., Prezzi, M., Salgado, R. & Zaheer, M. (2020).
of the offshore technology conference, Houston, TX, USA, 4, Axial resistance of open-ended pipe pile driven in gravelly sand.
pp. 2570–2587. Géotechnique 70, No. 2, 138–152, https://doi.org/10.1680/jgeot.
Arany, L., Bhattacharya, S., Macdonald, J. & Hogan, S. J. (2017). 18.P.117.
Design of monopiles for offshore wind turbines in 10 steps. Soil Klinkvort, R. T. & Hededal, O. (2013). Lateral response of monopile
Dyn. Earthq. Engng 92, 126–152. supporting an offshore wind turbine. Proc. Instn Civ. Engrs
Burd, H. J., Beuckelaers, W. J. A. P., Byrne, B. W., Gavin, K. G., Geotech. Engng 166, No. 2, 147–158, https://doi.org/10.
Houlsby, G. T., Igoe, D. J. P., Jardine, R. J., Martin, C. M., 1680/geng.12.00033.
McAdam, R. A., Muir Wood, A., Potts, D. M., Skov Kuwano, R. (1999). The stiffness and yielding anisotropy of
Gretlund, J., Taborda, D. M. G. & Zdravković, L. (2020a). sand. PhD thesis, Imperial College, University of London,
New data analysis methods for instrumented medium-scale London, UK.
monopile field tests. Géotechnique 70, No. 11, 961–969, LeBlanc, C., Houlsby, G. T. & Byrne, B. W. (2010). Response of stiff
https://doi.org/10.1680/jgeot.18.PISA.002 piles in sand to long-term cyclic lateral loading. Géotechnique
Burd, H. J., Taborda, D. M. G., Zdravković, L., Abadie, C. N., 60, No. 2, 79–90, https://doi.org/10.1680/geot.7.00196.
Byrne, B. W., Houlsby, G. T., Gavin, K. G., Igoe, D. J. P., Lehane, B. M. & Randolph, M. F. (2002). Evaluation of a minimum
Jardine, R. J., Martin, C. M., McAdam, R. A., Pedro, A. M. G. base resistance for driven pipe piles in siliceous sand. J. Geotech.
& Potts, D. M. (2020b). PISA design model for monopiles Geoenviron. Engng 128, No. 3, 198–205.
for offshore wind turbines: application to a marine sand. Loukidis, D. (2006). Advanced constitutive modeling of sands and
Géotechnique 70, No. 11, 1048–1066, https://doi.org/10. applications to foundation engineering. PhD thesis, Purdue
1680/jgeot.18.P.277. University, West Lafayette, IN, USA.
Byrne, B. W., Mcadam, R., Burd, H. J., Houlsby, G. T. & Loukidis, D. & Salgado, R. (2009). Modeling sand response using
Martin, C. M. (2015). New design methods for large diameter two-surface plasticity. Comput. Geotech. 36, No. 1–2, 166–186.
piles under lateral loading for offshore wind applications. In Luo, R., Yang, M. & Li, W. (2018). Numerical study of diameter
Frontiers in offshore geotechnics III (ed. V. Meyer), pp. 705–710. effect on accumulated deformation of laterally loaded
Abingdon, UK: Taylor & Francis. monopiles in sand. Eur. J. Environ. Civ. Engng 24, No. 14,
Byrne, B. W., Burd, H. J., Zdravkovic, L., Abadie, C. N., 2440–2452.
Houlsby, G. T., Jardine, R. J., Martin, C. M., McAdam, R. A., Byrne, B. W., Houlsby, G. T., Beuckelaers, W. J. A.
McAdam, R. A., Pacheco Andrade, M., Pedro, A. M. G., P., Burd, H. J., Gavin, K. G., Igoe, D. J. P., Jardine, R. J.,
Potts, D. M. & Taborda, D. M. G. (2019a). PISA design Martin, C. M., Muir Wood, A., Potts, D. M., Skov Gretlund, J.,
methods for offshore wind turbine monopiles. Proceedings of the Taborda, D. M. G. & Zdravković, L. (2020). Monotonic
offshore technology conference, Houston, TX, USA, paper laterally loaded pile testing in a dense marine sand at
OTC-29373-MS. Dunkirk. Géotechnique 70, No. 11, 986–998, https://doi.org/10.
Byrne, B. W., Burd, H. J., Zdravković, L., McAdam, R. A., 1680/jgeot.18.PISA.004.
Taborda, D. M. G., Houlsby, G. T., Jardine, R. J., Paik, K., Salgado, R., Lee, J. & Kim, B. (2003). Behavior of open-
Martin, C. M., Potts, D. M. & Gavin, K. G. (2019b). PISA: and closed-ended piles driven into sands. J. Geotech. Geoenviron.
new design methods for offshore wind turbine monopiles. Engng 129, No. 4, 296–306.
Rev. Fr. Géotech. 158, article 3, https://doi.org/10.1051/ Poulos, H. G. & Davis, E. H. (1980). Pile foundation analysis and
geotech/2019009. design. Hoboken, NJ, USA: John Wiley & Sons.
Chow, F. C. (1997). Investigations into the behaviour of displacement Purdue University (2021). https://script.google.com/macros/s/
piles for offshore foundations. PhD thesis, Imperial College, AKfycbwveAcQBPbghcXK9WHrktAFisBe8OfDIRvU-WfP--
University of London, London, UK. nLwT60XXi3i923u/exec (accessed 26/08/2021).
Cox, W. R., Reese, L. C. & Grubbs, B. R. (1974). Field testing of Randolph, M. F. (1981). The response of flexible piles to lateral
laterally loaded piles in sand. Proceedings of 6th offshore loading. Géotechnique 31, No. 2, 247–259, https://doi.org/10.
technology conference, Houston, TX, USA, pp. 459–472. 1680/geot.1981.31.2.247.
Davies, T. G. & Banerjee, P. K. (1978). The behaviour of axially and Randolph, M. & Gourvenec, S. (2017). Offshore geotechnical
laterally loaded single piles embedded in nonhomogeneous soils. engineering. Abingdon, UK: Taylor & Francis.
Géotechnique 28, No. 3, 309–326, https://doi.org/10.1680/geot. Reese, L. C., Cox, W. R. & Koop, F. D. (1974). Analysis of laterally
1978.28.3.309. loaded piles in sand. Proceedings of the offshore technology
DNV GL (2016). DNVGL-ST-0126: Support structures for wind conference, Houston, TX, USA, pp. 473–483.
turbines. Oslo, Norway: DNV GL. Reese, L. C., Cox, W. R. & Koop, F. D. (1975). Field testing and
Doherty, P. & Gavin, K. (2012). Laterally loaded monopile design analysis of laterally loaded piles in stiff clay. Proceedings of the
for offshore wind farms. Proc. Instn Civ. Engrs Energy 165, offshore technology conference. Houston, TX, USA, pp. 671–675.
No. 1, 7–17, https://doi.org/10.1680/ener.11.00003. Salgado, R. & Prezzi, M. (2007). Computation of cavity expansion
EIA (US Energy Information Administration) (2019). Electricity pressure and penetration resistance in sands. Int. J. Geomech. 7,
explained – electricity in the United States. Washington, No. 4, 251–265.
DC, USA: EIA. See https://www.eia.gov/energyexplained/ Salgado, R., Han, F. & Prezzi, M. (2017). Axial resistance of non-
electricity/electricity-in-the-us.php (accessed 13/09/2021). displacement piles and pile groups in sand. Rivista Italiana di
Fleming, K., Weltman, A., Randolph, M. & Elson, K. (2008). Piling Geotecnica 51, No. 4, 35–46.
engineering. Abingdon, UK: Taylor & Francis. Taborda, D. M. G., Zdravković, L., Potts, D. M., Burd, H. J.,
GWEC (Global Wind Energy Council) (2018). Global wind report Byrne, B. W., Gavin, K. G., Houlsby, G. T., Jardine, R. J.,
2018. Brussels, Belgium: GWEC. Liu, T., Martin, C. M. & McAdam, R. A. (2020). Finite-element
Han, F., Salgado, R., Prezzi, M. & Lim, J. (2017). Shaft and base modelling of laterally loaded piles in a dense marine sand at
resistance of non-displacement piles in sand. Comput. Geotech. Dunkirk. Géotechnique 70, No. 11, 1014–1029, https://doi.
83, 184–197. org/10.1680/jgeot.18.PISA.006.

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.
1050 HU, HAN, PREZZI, SALGADO AND ZHAO
Zdravković, L., Taborda, D., Potts, D., Jardine, R., Sideri, M., Zdravković, L., Jardine, R. J., Taborda, D. M. G., Abadias, D.,
Schroeder, F., Byrne, B., McAdam, R., Burd, H., Houlsby, G., Burd, H. J., Byrne, B. W., Gavin, K. G., Houlsby, G. T.,
Martin, C., Gavin, K., Doherty, P., Igoe, D., Wood, A., Igoe, D. J. P., Liu, T., Martin, C. M., McAdam, R. A., Muir
Kallehave, D. & Gretlund, J. (2015). Numerical modelling of Wood, A., Potts, D. M., Skov Gretlund, J. & Ushev, E. (2020).
large diameter piles under lateral loading for offshore wind Ground characterisation for PISA pile testing and analysis.
applications. In Frontiers in offshore geotechnics III (ed. V. Meyer), Géotechnique 70, No. 11, 945–960, https://doi.org/10.1680/jgeot.
pp. 759–764. Abingdon, UK: Taylor & Francis. 18.PISA.001.

Downloaded by [ Indian Institute of Technology Roorkee] on [18/04/23]. Copyright © ICE Publishing, all rights reserved.

You might also like