You are on page 1of 21

Numerical Investigation of Large-Diameter Monopiles in

Sands: Critical Review and Evaluation of Both API and Newly


Proposed p-y Curves
Djillali Amar Bouzid1
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Despite the increase in offshore wind turbine (OWT) capacities, the preferred foundations for these marine structures are large-
diameter monopiles due to their ease of installation in shallow to medium water depth. The current design methodologies based on p-y curves
[the American Petroleum Institute (API) method, for instance] have gained a confirmed recognition in designing slender monopiles, such as
those used for supporting offshore gas and oil platforms. However, when applied to large-diameter monopiles, these regulation codes failed to
properly address the behavior of these extremely stiff structures, which is why many researchers tried to enhance the p-y method performance
by suggesting new p-y formulas incorporating monopile properties. The paper reviews the analytical formulas of the p-y curves adopted by the
current offshore guidelines and documents in detail the limitations and the shortcomings inherent to their formulations. The latest versions of
p-y curves proposed to improve the performance of the Winkler model were also given. To show how well the finite-element (FE) analysis pre-
dicts the lateral response of large-diameter monopiles under horizontal loading on one hand, and how inappropriate the models using p-y
curves are in studying this crucial problem on the other hand, two Fortran computer programs were considered. The first program, a computer
code called NAMPULAL, combines the FE vertical slices method (VSM) and the hyperbolic model to describe the behavior of cohesionless
soils. The second program, called Winkler-ROWKSS, is written using the finite-differences method. In the latter, the original and the newly
developed p-y curves were implemented. The comparative study was made by considering two case histories. The first case is a monopile em-
bedded in a multilayer sandy soil, which serves as a support foundation to 5-MW National Renewable Energy Laboratory (NREL) wind tur-
bines, and the second case is a monopile supporting OWTs at Horns Rev in the Danish sector of the North Sea. At the first site, the evolution
of the monopile head movements with applied loading provided by the different models of beams on nonlinear Winkler foundations
(BNWFs) as well as the FEM was examined and compared. However, at the second site, only the monopile lateral displacement profile
was studied. DOI: 10.1061/(ASCE)GM.1943-5622.0001204. © 2018 American Society of Civil Engineers.
Author keywords: Winkler model; p-y curves; American Petroleum Institute (API); Finite-element vertical slices model (FE VSM);
Large-diameter monopiles; Offshore wind turbines (OWTs).

Introduction shallow depths up to 40 m because they are cost-effective and rela-


tively simple to install (Ahmed and Hawlader 2016; Abed et al.
The depletion of traditional energy sources has led to an increase in 2016; Aissa et al. 2017). The monopiles consist of welded steel pipe
the search for alternatives in renewable energy sources to supply piles driven open ended into soil. Monopiles with diameters up to
societal needs. Consequently, there has been rapid development in 6 m are already installed and even larger diameters are currently
the installation of wind farms in both onshore and offshore locations planned (Achmus et al. 2009; Adhikari and Bhattacharya 2011,
to secure clean electricity as a permanent source of energy. Because 2012; Carswell et al. 2015; Lombardi et al. 2013; Arany et al. 2016,
offshore winds tend to have higher speeds and are less turbulent 2017; Galvín et al. 2017; Sheil and Finnegan 2017).
than onshore winds, resulting in a greater increase in the potential The key element in the design of monopile foundations is to ful-
energy produced in offshore locations than in onshore ones, several fill the ultimate limit state (ULS) and the serviceability limit state
offshore wind farms have been already installed all over the world (SLS) design analyses. In the ULS analysis, the monopile properties
in the last decade, and a vast number of farms are planned for the should be chosen so that soil resistance is kept within a safe margin
near future. Although there are many offshore wind turbines far away from the soil failure to ensure the structure safety of the
(OWTs) to support options depending on water depth (ranging from entire OWT. In the SLS proof, the monopile head movements
gravity foundations for shallow depths of 0–15 m to floating foun- (deflection and rotation) under extreme load cases have to remain
dations for very deep waters of 60–200 m), the monopile is the pre- below certain limits.
ferred foundation concept for offshore wind energy converters in To satisfy both ULS and SLS requirements, the design of the
structure and the substructure of an OWT should be performed
1
properly by including the soil/monopile interaction. This is done so
Professor, Dept. Civil Engineering, Faculty of Technology, Univ. that OWTs sustain the permanent dynamic forces induced by vibra-
Saad Dahled of Blida, Route de Soumaa, Blida 09000, Algeria. Email:
tions during their operational life because they are dynamically sen-
d_amarbouzid@yahoo.fr
Note. This manuscript was submitted on September 13, 2017;
sitive structures. The induced forces combined with the operating
approved on February 2, 2018; published online on August 23, 2018. frequency could potentially trigger the resonance phenomenon.
Discussion period open until January 23, 2019; separate discussions must To avoid this problematic phenomenon, which may lead to the
be submitted for individual papers. This paper is part of the International entire damage of an OWT, the latter should be designed such that
Journal of Geomechanics, © ASCE, ISSN 1532-3641. the first eigenfrequency is sufficiently far away from the main

© ASCE 04018141-1 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


excitation frequencies of the dynamic loading. In other words, it Murchison (1983) formulation, which was adopted first by the API
should lie between turbine (1 P) and blade-passing frequencies and then by DNV, is no longer appropriate for the analysis of large-
(3 P), i.e., in the interval corresponding to the “soft-stiff” interval. diameter monopiles under a horizontal loading setting thus, far
This permissible interval frequency is shown in the Campbell dia- apart from the FE analysis. In other words, the Winkler models
gram of Fig. 1 (Bisoi and Haldar 2015). inappropriately predict the response of laterally loaded monopiles.
The American Petroleum Institute (API 2014) and Det Norske The third objective is to properly evaluate the Winkler models based
Veritas (DNV 2013) standards, which are the most used offshore on the latest proposed modifications of sand p-y curves in the range
guidelines, recommend using the Winkler model approach to bounded by both API and FE results. Here, the purpose is to get a
design monopiles subjected to lateral loading. In this approach the better overview of whether these models have improved the accurate
monopiles are considered as beams on nonlinear Winkler founda- assessment of the lateral behavior of large-diameter monopiles sup-
tions (BNWFs), and the soil response is analyzed by means of p-y porting OWTs or failed to correctly address this issue.
curves providing the soil reaction p in the function of lateral deflec- To prepare the necessary material to achieve these distinct objec-
tion y. However, this method of analysis fails when applied to large- tives and for a better understanding of the behaviors of both small-
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

diameter monopiles (with diameters of 5–8 m and slenderness ratios and large-diameter monopiles, the next section is dedicated to a
around 5). The limitations of this empirical method are discussed in description of the failure mechanisms of these kinds of deep foun-
detail in this paper in a later section. dations followed by the p-y formulations adopted by the offshore
As far as fulfilling both ULS and SLS requirements is concerned, guidelines from API and DNV. In this section the author provides
the monopile design should rely on a rigorous method, such as the in detail a critical view of BNWF as a method of analysis and gives
FEM. In this regard, the computer code called nonlinear analysis of the limitations of its application to large-diameter monopiles. In a
monopiles under lateral and axial loading (NAMPULAL), written subsequent section the most recently proposed modifications of the
on the basis of the finite-element (FE) vertical slices model (VSM) sand p-y curves are given. These modified p-y curves have been
combined with the use of a hyperbolic model for modeling sandy implemented by the author of this paper in a computer program
soils (Otsmane and Amar Bouzid 2018), is used to analyze the called Winkler-ROWKSS to be used later for comparison purposes.
monopiles under monotonic loading. The three objectives targeted by the present work are achieved by
The objective of this paper is divided into three separate subtar- the analysis of two case histories.
gets. The first objective is to show that an FE analysis based on a
nonlinear soil hyperbolic model implemented in NAMPULAL is
the best choice as a method of analysis for the safe design of OWTs. Flexible and Stiff Failure Mechanisms of Small- and
This is achieved by examining the monopile head displacements Large-Diameter Monopiles
and rotations by NAMPULAL as well as the monopile displace-
ment profile, which is consistent with the deformation pattern of Single piles (monopiles) are those used to support high-rise build-
stiff monopiles. The second objective is to confirm statements about ings, bridges, or offshore platforms in the oil and gas industry. They
the behavior of laterally loaded monopiles. Since the emergence of can be distinguished in terms of lateral behavior and categorized
the OWT industry, many researchers ascertain that the O’Neill and into two main classes, namely flexible (slender) piles and short

0 .6

on
a ti
c it
0 .5 ex
S t iff- S t if f 3P
S o ft - S t iff

0 .4
Frequ en cy (H z)

A c c e p t a b le r a n g e
o f fr e q u e n c y
0 .3 ta tio
n
xci
1P e

0 .2

S o ft - S o ft

0 .1

0 .0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17

R o to r sp e e d (r p m )

Fig. 1. Campbell diagram. (Data from Bisoi and Haldar 2015.)

© ASCE 04018141-2 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


(rigid) piles. In the first class, the behavior of monopiles is charac- certain OWT dynamic characteristics depending to a large extent on
terized by the movement of only the upper part, whose length the nature of the monopile. In long flexible monopiles, the lateral
depends on soil/monopile stiffness. In the extreme situation, the spring stiffness KLFlexible , the rocking spring stiffness KRFlexible , and
Flexible
failure occurs in the pile because of the formation of plastic hinges the cross-coupling spring stiffness KLR depend, in addition to
at zero displacement locations. In the second class, a “toe kick” soil Young’s modulus Es and monopile diameter Dp , on the
occurs in response to the applied loads; in other words, the entire monopile/soil relative rigidity Ep =Es . Appendix I gives the most
monopile length undergoes movement. The failure mechanism for known established relationships. However, for the large-diameter
stiff monopiles occurs in the soil first as the monopile undergoes a monopiles and the small-diameter short monopiles, the stiff-
significant rotation at its head. Because the diameter is generally nesses KLRigid ; KRRigid , and KLR
Rigid
depend, in addition to modulus Es
less than 2 m, this flexibility or rigidity depends on whether the and monopile diameter Dp , on the monopile slenderness ratio
monopile’s effective length Lp is less than or greater than a value Lp =Dp . This was stated more than two decades ago by Carter and
termed critical length Lpc [Figs. 2(b and c)]. These small-diameter Kulhawy (1992) for homogeneous soils and confirmed recently by
piles are short and stiff if their effective length is less than the criti- many authors (Higgins et al. 2013; Shadlou and Bhattacharya 2016;
cal length, and slender and flexible in the reverse situation. The con-
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

Abed et al. 2016; Aissa et al. 2017) for three different soil profiles.
cept of critical length or critical slenderness Lpc =Dp (Dp = pile diam- These stiffness impedance functions are listed in Appendix I.
eter) is the main distinguishing key between these two behaviors.
The notion of critical length has been well addressed by many
researchers, decades ago (Randolph 1981; Davies and Budhu 1986; P-Y Formulations as Adopted by API and DNV
Amar Bouzid 1997).
Although monopiles used to support OWTs are deeply driven In current geotechnical practice, analysis of laterally loaded piles
into marine subsoils, in which the embedment length may reach is commonly conducted using an approach in which the soil-
35 m in some situations, their effective length is less than the corre- foundation system is modeled as a BNWF with the soil repre-
sponding critical length because of their large diameter, which is sented by a series of uncoupled springs whose response is
currently 6 m in routine practice and planned to be 8 m for future described by a curve called the p-y curve (Fig. 3). This model
installations (Thieken et al. 2015). Consequently, these monopiles involves the resolution of the fourth-order differential equation
are long and stiff. Fig. 2(a) shows the mechanism of failure of a d4 y
monopile, which takes place by excessive monopile top and bottom ðEI Þp ðzÞ p¼0 (1)
dz4
rotations, and there is no plastic hinge formation.
The stiffnesses characterizing the springs are usually used to rep- where y = lateral deflection of the monopile at a point z along the
resent the soil/monopile interaction; hence, it is used to compute monopile length; p = soil reaction in force unit per unit length; and

(a) Tilting of an OWT founded on a monopile

(b) Tilting of a short monopile in


bridge engineering or building
sectors

(c) Deflection of small diameter long (e) Springs representing


(d) Springs representing the SSI the SSI in (c)
monopiles experiencing two or more
in (a) and (b)
points of zero deflection.

Fig. 2. Failure mechanisms of large- and small-diameter monopiles used to support both OWTs and bridges, respectively, in energy and building sec-
tors. SSI = soil–structure interaction.

© ASCE 04018141-3 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


Ficve monopile +1
overhang ℎ

−1 Imaginary nodes
at the monopile
Spring −1 −1
−1
−2 ℎ head

Spring −2 −2
−2

−2

Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.


Spring

+1

+2
Imaginary monopile
p segment 2

Spring 1 1 Imaginary nodes


ℎ at the monopile p
0


−1

−2

(a) (b) (c) (d)

Fig. 3. BNWF model applied to a slender monopile: (a) under lateral loading; (b) spring model; (c) prescribed p-y curves; and (d) finite-difference
modeling of the monopile.

ðEI Þp ðzÞ = monopile flexural rigidity, which may vary with depth, piles installed at Mustang Island, Texas. A total of seven load
as in the case of segmented or tapered monopiles. tests were performed on the piles consisting of two static tests
Over six decades, the solution of Eq. (1) has been usually and five cyclic tests. Both piles had a diameter of 0:61 m and a
based on the combination of two fundamental concepts. The first, slenderness ratio of 34:4, and both were installed in a medium
relevant to the lateral behavior of the monopile, is the use of the sand. Based on these field tests, Reese et al. (1974) proposed a
finite-differences method to discretize the monopile into n seg- semiempirical p-y curve consisting of four segments. Using this
ments resulting in ðn þ 1Þ nodes. Because Eq. (1) is a fourth-order piecewise p-y curve proposed by Reese et al. (1974) and a rela-
differential equation, the formulation of finite differences at the tively large database of laterally loaded pile tests, O’Neill and
level of each node results in a five-term equation containing the Murchison (1983) suggested a hyperbolic formula for the p-y
lateral displacement of the targeted node and the displacements of curve to describe the relationship between soil resistance and
the four adjacent nodes. Consequently, two imaginary points over lateral pile deflection in sand
the monopile head and two imaginary points below the monopile
 
tip are required to find yn and y0 , respectively (Fig. 3). The entire kz
monopile finite-difference modeling results in ðn þ 5Þ unknown pðy; zÞ ¼ Apu tanh y (2)
Apu
lateral deflections. The second concept is the use of p-y curves,
which are provided at each node level giving the soil response in
where k represents the initial coefficient of subgrade reaction
terms of lateral deflections. The global solution depends funda-
depending on the angle of internal friction or the relative density of
mentally on the assumption made in the equations encompassing
the cohesionless soil. Its values can be obtained either from Fig. 4(a)
the imaginary nodes, especially at the monopile tip. The solution
or from Eq. (3), which has been fitted from Fig. 4(a) by Augustesen
proceeds iteratively using both concepts, in which the soil reac-
et al. (2009)
tion in the current iteration is determined on the basis of the
deflection found in the previous iteration, until a convergence cri-  
terion is fulfilled. k ¼ 0:008085 f 2:45  26:09 103 ½kPa=m for 29  f  45
The original formulation of p-y curves goes back to the early (3)
1970s, in which there has been a great demand for designing piles
supporting platforms in the gas and oil offshore industry for where f is in degrees.
which the soil–pile interaction became crucial to analyze. In that The ultimate resistance pu is determined by the following
period Reese et al. (1974) tested two identical instrumented equation:

© ASCE 04018141-4 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


'
φ , Angle of internal friction
nn
100
30°
28° 29° 36° 40° 45°

Very Medium Very


Loose Dense
Loose Dense Dense
80

Sand above
60 the water table

40
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

Sand below
20 the water table
k (MPa/m)

0
0 20 40 60 80 100
(a) Relative density Dr (%)

5 100

90

4 80

Values of coefficients c3
70
Values of coefficients c1 , c2

3 60

50

2 40

C3 30

1 C1 20
C2
10

0 0
20 25 30 35 40
'
Angle of internal friction φ (°)
(b)

Fig. 4. (a)Variation of initial coefficient of subgrade reaction k as a function of relative density; and (b) variation of the coefficients C1 , C2 , and C3 as
a function of the angle of internal friction. (Data from DNV 2013.)

8
< ðC1 z þ C2 Dp Þ g z The coefficient A is determined from the equations
pu ðzÞ ¼ min (4)  
: ðC3 Dp g zÞ z
A ¼ 3:0  0:8  0:9 for static loading (6)
Dp

where the coefficients C1 , C2 , and C3 are functions of the angle of A ¼ 0:9 for cyclic loading (7)
internal friction, which can be either determined using the graph of
Fig. 4(b), or the following expressions: The original p-y curve proposed by Reese et al. (1974) has been
adopted by the API for designing piles supporting platforms in the
C1 ¼ 0:115 100:0405 f ; C2 ¼ 0:571 100:022 f ; and oil and gas industry and then was replaced later by the hyperbolic
formula of Eq. (2). This formula is still in use in the current design
C3 ¼ 0:646 100:0555 f (5) practice of monopiles.

© ASCE 04018141-5 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


Critical Review of the p-y Curve Method Loading
Because many inconsistencies and drawbacks have been observed ,
in the previous p-y formulations when dealing with monopiles
under horizontal loading, the critical review presented in this sec- ≤ 30 for
tion is organized according to three broad categories, and each is most
thoroughly detailed. monopiles Wedge
installed so far failure
Drawbacks in the Original p-y Formulation as Proposed
by Reese et al. (1974)
The p-y curve suggested by Reese et al. (1974) consists of three Horizontal flow
straight lines and a parabola. The limiting lateral displacements sep- around
arating the segments have been set empirically without any theoreti-
cal evidence, although the entire method has been developed on the Cohesionless soil
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

basis of a theoretical background. From the considerable amount of = 0, ≠ 0.


empiricism that has been involved in the recommendations by
Reese et al. (1974), two significant limitations are worth detailing in
this paper.
Limitation Relevant to the Earth Pressure Coefficient at Rest. Fig. 5. Soil deformation models according to Reese et al. (1974).
Among the soil parameters that affect the soil behavior is the earth (Data from Reese et al. 1974.)
pressure coefficient at rest K0 . Although simple, this coefficient is
by far the most influential parameter with respect to the lateral
behavior of monopiles. In fact, it has been overlooked by research-
ers in the original formulation of Reese et al. (1974), who adopted a Sand friction angle φ (°)
30 32 34 36 38 40 42
value of 0:4 for K0 for the calculation of the soil reaction at failure. 0
This meant that in Jâky’s (1944) point of view, in which he pro-
posed K0 ¼ 1  sin f , Reese et al. (1974) developed their theory for 20
a sand whose friction angle is equal to f ¼ 37 , and generated the
p-y curves for any sand, regardless of its density state. Researchers 40

who investigated this parameter concluded that the earth pressure


60
coefficient at rest is dependent on many soil physical parameters,
such as the void ratio, the friction angle, the relative density, and the 80
overconsolidation ratio (OCR), for cohesive soils as well as the pro-
cess by which the soil deposit was formed (Jâky 1944; Mayne and 100
Kulhawy 1982; Michalowski 2005; Sherif et al. 1984). The values
Depth ZWF (m)

Dp=0.61 m
of K0 vary significantly when the sand varies from a loose state to a 120
Dp=2 m
very dense state. Consequently, the most affected parameter is the Dp=4 m
pile displacement and therefore the ultimate soil resistance. In this 140
Dp=6 m
context, and using a computer code based on a conventional three-
160
dimensional (3D) FEM, Dickin and Laman (2003) analyzed a short
pier in sandy soil and examined the variation of the pier head rotation Fig. 6. Variation of ZWF in function of f and Dp .
in the function of the variation of K0 . They found that the pier head
rotations are markedly sensitive to the values of K0 , especially those
ranging between 0:25 and 0:60, and relatively minor variations were two concepts are equal. The author of this paper performed the cal-
noticed in rotations for K0 values greater than 0:60. Fan and Long culations and found this depth as
(2005) performed FE analyses on laterally loaded piles and studied
the effect of varying K0 on the shape of p-y curves. They found not a5 a2 a
only an increase in the initial p-y curve stiffness but also a significant þ 2k0 4 
b5 ð1 bÞ b bð1  bÞ
increase in the ultimate soil resistance. ZWF ¼ Dp   (8)
2k0 a2 2d
Ambiguous Definition of the Depth at Which the Ultimate Soil pffiffiffi þ 2 þ k0 pffiffiffi  1
Resistance Changes from Wedge Type to Flow-Around Type. 2b b 2b
Reese et al. (1974) adopted two soil models for computing the ulti-
mate soil resistance. Using a wedge model near the ground surface where a ¼ 1 þ tanð f =2Þ, b ¼ 1  tanð f =2Þ, and d ¼ cosð f =2Þ;
(at shallow depth) and on the basis of the Rankine theory and the Dp = pile diameter; and f = internal friction angle of sand.
Mohr-Coulomb failure condition, they formulated an expression for Calculation steps leading to Eq. (8) have been detailed in
the ultimate soil resistance containing the pile diameter, the depth Appendix II.
of the sand wedge, and most importantly the sand friction angle. At In the author’s opinion, for large-diameter monopiles, ZWF is
some distance from the ground level, which is not quantified, they simply a depth at which the two failure concepts yield the same
considered another model of sand failure that consists in a horizon- value of ultimate pressure occurring well beneath the monopile tip.
tal flow around the pile. A different expression for the ultimate re- It does not define a shallow depth of a few diameters at which the
sistance has been proposed containing the same parameters (Fig. 5). ultimate sand resistance changes from the first soil model to the sec-
The shift in the failure mechanism from the first model to the second ond one.
has been set to occur according to Reese et al. (1974) at a particular To discuss this issue, the variation of ZWF in the function of the
depth ZWF (Fig. 5) in which the ultimate pressures provided by the sand friction angle f for a number of monopile diameters has been

© ASCE 04018141-6 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


plotted for examination in Fig. 6. The figure shows that even for a medium elastic properties, Vesic (1961) demonstrated that the mod-
diameter Dp ¼ 2 m, which most authors claim as a limiting diame- ulus of subgrade reaction is independent of the pile diameter
ter of Winkler model validation, the entire monopile embedment, because the latter disappears due to the fact that it has the same
which is in most cases less than 30 m, falls within the shallow depth power in both the numerator and denominator of Eq. (10)
concept according to Reese et al. (1974). For monopiles with larger " #1=2
diameters the shallow depth is even deeper. Consequently, the con- 0:65Es Es D4p
cept of soil failure as a horizontal flow around the monopile would K¼  (10)
1  s2 ðEI Þp
never take place because all monopiles are larger than 2 m. A more
appropriate definition should be considered using a different theo-
retical background. where, Es = soil modulus of elasticity; s = soil Poisson’s ratio;
Dp = monopile diameter; and EIÞp = monopile flexural rigidity.
Shortcomings in the p-y Formulation as Proposed by Pender et al. (2007) and Ashford and Juirnarongrit (2003) con-
O’Neill and Murchison (1983) ducted a series of FE analyses to investigate the effect of the monop-
ile diameter on the initial stiffness of the p-y curve. The research,
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

The analytical formula [Eq. (2)] proposed by these authors and


adopted by the API regulation code was attractive because it is sim- however, concluded that there is a negligible impact of the pile diam-
ple to implement in computer code. However, the formula failed to eter. Fan and Long (2005) also performed a FE investigation in
design large-diameter monopiles in the same way as the formulation which they kept the monopile bending stiffness ðEI Þp constant and
by Reese et al. (1974) did. The major drawbacks in the formulation increased the monopile diameter. This research confirmed the same
by O’Neill and Murchison (1983) lie in the fact that the initial stiff- observation. However, the numerical modeling of the monopiles
ness of their p-y curve significantly overestimates the soil stiffness under lateral loading by Wiemann et al. (2004) and Sørensen et al.
at large depths and does not include any monopile geometrical (2010) reported that the initial stiffness of the soil–monopile interac-
property. This has been considered by many researchers to be at the tion is affected by changing the monopile diameter.
origin of the poor performance of this Winkler model. Although significant, the influence of pile bending stiffness on
Monopile Geometrical and Bending Stiffness Properties Are the p-y curve has not been given much attention in the literature.
Disregarded. The initial stiffness of the p-y curves Epy 
recom- Only Ashour and Norris (2000) and Fan and Long (2005) have
mended in the design can be obtained by evaluating the slope of the focused on this issue. Adopting the strain wedge method, the first
p-y curve tangent at y ¼ 0 investigation confirmed the effect of ðEI Þp on the soil ultimate re-
sistance, whereas the second investigation reported no significant
kz influence of ðEI Þp on the p-y curve.
 
dp  Apu  In addition to the confused situation of whether the p-y constitu-

Epy ¼  ¼ Apu    ¼ kz (9) tive elements k and pu depend on the monopile diameter and its
dy y¼0 kzy
cosh2 y¼0
bending stiffness, it is also difficult to find k for a very dense sand
Apu
that has an internal friction angle greater than 40 [Fig. 4(a)]. As a
consequence of these inconsistencies, many researchers reported
It is quite clear from Eq. (9) that the initial stiffness is independ- that the linear variation of stiffness with depth, as shown by Eq. (9),
ent of pile properties (diameter and bending stiffness), and it is line- leads to an overestimation of soil stiffness at large depths.

arly dependent of the depth z. Because Epy , through the parameter k, Furthermore, the currently used design codes lead to an underesti-
relies only on the sand relative density and does not depend on the mation in the head displacements for large-diameter monopiles
monopile properties, many questions can arise because research on under extreme loads (Abdel-Rahman and Achmus 2005; Wiemann
this issue gave contradictory conclusions. We will see in the re- et al. 2004; Sørensen 2012) and to an overestimation for monopiles
mainder of this paper that most authors who tried to enhance the under small operational loads (Kallehave et al. 2012; Hald et al.
performance of the p-y curves intervened at the level of the initial 2009). These facts have been recently confirmed by Otsmane and
stiffness of the p-y curve rather than at the level of the ultimate pres- Amar Bouzid (2018).
sure. The literature about this issue is scarce, and very little research These contradictory conclusions are summarized chronologi-
has been performed over nearly six decades. Indeed, Terzaghi cally in Table 1.
(1955) provided an explanation of how the depth of the stress bulb
can be affected by the pile diameter whether it is small or large. He Drawbacks in the BNWF as a Method of Analysis
concluded that the modulus of subgrade reaction is independent of of Large-Diameter Monopiles
the pile diameter. By providing the expression given by Eq. (10), The poor performance noticed in the BNWF approach when
which relates the modulus of subgrade reaction to both pile and applied to large-diameter monopiles is due to the extrapolation of

Table 1. Conclusions reached by some researchers about the effects of Dp and ðEI Þp on the p-y curve

Effect of the monopile Effect of the monopile


Authors diameter Dp bending stiffness ð EI Þp Method used
Terzaghi (1955) No effect — Full-scale experiments
Vesic (1961) No effect — Empirical research
Ashour and Norris (2000) — Important effect Numerical analysis
Ashford and Juirnarongrit (2003) Negligible effect — Experimental program including vibration and impact tests
Wiemann et al. (2004) Significant effect — 3D FE analysis
Fan and Long (2005) Negligible influence No significant effect 3D FE analysis
Pender et al. (2007) Apparent effect — Numerical computations
Sørensen et al. (2010) Confirmed effect — Numerical computations using FLAC3D

© ASCE 04018141-7 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


p-y curves to this kind of foundation (originally calibrated for monopile. In this regard, Ibsen et al. (2009) proposed an empirical
small-diameter piles) and to the BNWF model itself as a method equation giving f in the function of sand relative density Dr as well
of analysis. Shortcomings inherent to this Winkler model can be as the minor principal stress s 3
summarized as follows.
Inconsistency between the Monopile Behavior Pattern and f ¼ 15 Dr þ 27 s 30:28 þ 23½  (13)
Boundary Conditions Adopted. Regarding the behavior of either
short monopiles used in the sector of bridge engineering or in the oil This equation cannot be introduced in the Winkler model
and gas industry, or large-diameter monopiles used to support wind regardless of the quality of the p-y curve; however, it can be easily
energy converters, there is only one point of zero deflection along introduced into an FE analysis. Eq. (13) confirms that the sand
the monopile length. In this case, the deflection of the bottom of the stiffness is governed to a large extent by the confining pressure
monopile could cause a shearing force to develop. However, the s 3 , showing the suitability of hyperbolic models for the descrip-
finite-difference procedure generally formulated to solve the fourth- tion of sand behaviors as it constitutes the key element in their
order differential [Eq. (1)] does not distinguish between flexible and formulations. This kind of model is used in the FE analysis pre-
stiff monopiles. This is supported by the fact that most written
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

sented subsequently in this paper, and Eq. (13) is implemented to


Winkler approach-based codes, such as COM624 (Reese and follow the sand stiffness variation during the process of loading/
Sullivan 1980) or LPILE6 (Wang and Isenhower 2010), have been unloading.
established on the assumption that the monopile will experience
two or more points of zero deflection (Reese and Van Impe 2011). Design Parameters Are Inaccurately Estimated in the BNWF
Thus, the hypothesis of the zero shearing force and zero bending In the BNWF model the design parameters, such as rotations, bend-
moment at the bottom of the monopile is usually adopted to find the ing moment, shear forces, and earth pressures, along the monopile
complete set of unknowns. However, because monopiles exhibit shaft are inaccurately computed because they are all related to the
one zero deflection along their entire length, the deflection at the lateral displacement by a differential equation with increasing
bottom could give birth to a shear force. In these circumstances, and order. In this model the obtained nodal displacements are used in a
referring to Fig. 3(d), the needed equations of boundary conditions finite-difference scheme to evaluate the previously mentioned pa-
encompassing imaginary points at the bottom of monopile are rameters. In this regard, it is well known that the finite-difference
method loses its accuracy when the differential equation order gets
R0
ðy2  2y1 þ 2y1  y2 Þ ¼ V0 (11) bigger. However, in the FEM only the monopile rotations are based
2h3 on the finite-difference method, and the shear and normal stresses
that develop at the vicinity of the monopile are integrated over the
R0 monopile circumference and along the monopile shaft to evaluate
ðy1  2y0 þ y1 Þ ¼ M0 (12)
h2 the other design parameters. The results are more accurate if the
monopile/soil interface is properly modeled (Amar Bouzid et al.
where R0 = monopile bending stiffness at node 0. It is difficult to 2004).
find appropriate values for V0 and M0 without any additional infor-
mation (Reese and Van Impe 2011). This makes the Winkler model
less viable in analyzing large-diameter monopiles and hampering Latest Propositions to Enhance the BNWF Model:
thus, to obtain accurate profiles of deflections along the monopile Computer Code Winkler-ROWKSS
shaft.
Because the standard p-y method provided an unsatisfactory basis
BNWF Cannot Update the Soil Stiffness during the for the design of wind turbine monopile foundations, many
Loading/Unloading Process researchers proposed analytical formulas for p-y curves in an
During the process of monopile loading, the sand can densify in attempt to enhance the analysis by accounting for the monopile di-
front of the monopile head and in the back of the monopile tip while ameter and sometimes for monopile stiffness parameters. Hence,
it loosens at the back of the monopile head and in front of the pile most of them suggested modifications altering the shapes of the p-y
tip. Consequently, the angle of internal friction f may change sig- curves at the level of their initial stiffness, whereas they considered
nificantly during this process in the vicinity of a large-diameter the formulation by Reese et al. (1974) adequate for the design of

Table 2. Proposed formulas for the p-y curve initial stiffness

Authors Proposed formulas Assigned parameters



O’Neill and Murchison (1983) (API 2014; DNV 2013) Epy ¼ kAPI z kAPI ¼ k appearing in Eqs. (2) and (3)
Wiemann et al. (2004) !4ð4þa
1aÞ
a ¼ 0:6 for a medium dense sand,
Dref

Epy ¼ kWiemann z ¼ z kAPI
p a ¼ 0:5 for a dense sand,
Dp
!c p ¼ 1:0 m
Dref
 b a ¼ 50;000; zref ¼ 1 m; Dref
Sørensen et al. (2010)
 z Dp p ¼ 1 m,
Epy ¼ kSørensen1 z ¼ a fd b ¼ 0:6; c ¼ 0:5; d ¼ 3:6, f in radians
zref Dref
p
 !0:5
m
Kallehave et al. (2012) Dpz m ¼ 0:6; zref ¼ 2:5 m,

Epy ¼ kKallehave z ¼ kAPI zref D ref
zref Dref
p p ¼ 0:61 m
 b !  d
Sørensen (2012) z Dp
c
Es a ¼ 1 MPa; zref ¼ 1 m; Dref
p ¼ 1 m,
Epy ¼ kSørensen2 z¼a Eref ¼ 1 MPa,
zref Dref
p Esref s
b ¼ 0:3; c ¼ 0:5; d ¼ 0:8

© ASCE 04018141-8 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


large-diameter monopiles and kept the ultimate soil resistance pu (2012) introduced a stress-level and diameter correction into the
unaltered in their new p-y curve formulations. API p-y curve. Unlike the other researchers, Kallehave et al. (2012)
Wiemann et al. (2004) noticed that the initial stiffness in the chose a reference monopile diameter, which is the diameter used by
original API approach overestimates grossly the soil stiffness at Reese et al. (1974) in their Mustang Island experiments. These dif-
large depths. Consequently, they proposed a formula that mitigates ferent formulas are listed in Table 2.
the soil stiffness by introducing a monopile diameter ratio and an The computer code Winkler-ROWKSS, which stands for
exponent less than one. Sørensen et al. (2010) conducted numerical Winkler model using Reese et al. (1974), O’Neill and Murchison
experiments using the commercial package FLAC3D in which they (1983), Wiemann et al. (2004), Kallehave et al. (2012), Sørensen
considered the monopile diameter and soil/monopile interaction et al. (2010), and Sørensen (2012) p-y curves, has been written by
effects on the initial stiffness of the proposed p-y curve. Further the author of this paper on the basis of an existing Fortran computer
modifications have been proposed by Sørensen (2012) in which soil code called COM622, which is still in use, to deal with axially and
Young’s modulus has been added and the sand friction angle horizontally loaded piles (Reese and Manoliu 1973; Reese 1977).
removed. On the basis of theoretical considerations, Kallehave et al. COM622 numerically solves Eq. (1) using the finite-differences
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

Medium to
be sliced

(a)

V The slice
M The irst three slices

3
2

2
3

2 3

(b)

Fig. 7. (a) OWT as a soil–structure interaction problem; and (b) the VSM showing the interacting slices subjected to external and body forces.

© ASCE 04018141-9 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


method. It is based on a special flowchart in which the different curves are provided analytically in both deflection (y) and depth
computation steps are performed iteratively, and the soil reaction in (z) values. This program is used for comparison purposes with FE
the current iteration is determined on the basis of the deflection analysis.
found in the previous iteration, until a convergence criterion is satis-
fied. The major drawbacks of this computer code are threefold.
First, the implementation is limited to deal with only experimental FE VSM Combined with the Hyperbolic Model for
p-y curves provided as input data at a very limited number of depth Modeling Cohesionless Soils
levels, which are in most cases insufficient to cover the entire
monopile length. Second, for a node that does not coincide with a The concept of the so-called VSM is fundamentally based on subdi-
p-y curve depth level, the computer interpolates between the two viding the structure and the surrounding soil into vertical panels of
sandwiching experimental p-y curves to compute the soil pressure. different thicknesses in the z-direction, in which two-dimensional
This pressure is less accurate if the distance between the two consid- (2D) conventional FEs are used to discretize each panel. The 3D as-
ered p-y curves is bigger. Third, because the soil reactions within pect of the problem is accounted for by coupling the shear forces
the experimental p-y curve are provided only at discrete points, the
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

between the slices. A soil/monopile interaction problem portrayed


computer code uses a linear interpolation to find the soil reaction in Fig. 7 shows the 3D problem [Fig. 7(a)] and the VSM in which
between two successive soil pressures in the same p-y curve. the different slices are acted on by external forces and body forces
To enable the calculation of the behavior of open pipe piles [Fig. 7(b)].
under lateral loading, the originally developed p-y curves and
those recently enhanced expressions that appeared in Table 2
have been implemented as separate subroutines for each model in Main FE Equations Describing the VSM
Winkler-ROWKSS. No interpolations are needed because the p-y The response of an individual slice to an external loading is gov-
erned by the following equilibrium equations:

∂s x ∂t xy ∂t xy ∂s y
þ þ bx ¼ 0; þ þ by ¼ 0 (14)
∂x ∂y ∂x ∂y
Deviator stress

Ulmate deviator stress ( − )


where bx and by are considered to be the key elements in the devel-
Deviator stress at failure ( − ) opment of the FE VSM. These physical parameters, termed body
Hyperbola
1
forces in this model, have been interpreted as fictitious forces trans-
mitted to the slice under consideration by shear forces applied from
Actual test slices at the left and at the right. For a given slice i, this can be math-
ematically expressed as

Hyperbolic representaon t lzxi  t rzxi t lzy  t rzyi


1 bx ¼ ; by ¼ i (15)
( − )= ti ti
+
( − )
where t lzxi and t lzyi are shear stresses acting at the left interface of
slice i; t rzxi and t rzyi are shear stresses acting at the right interface of
Axial strain ( ) slice i; and ti = slice thickness. The stress and deformation analyses
in each slice are conducted by the conventional FEM, using 2D
Fig. 8. Hyperbolic model of stress–strain behavior. FEs. According to the standard formulation in the displacement
based FEM, the element stiffness matrix in slice i can be written as

Table 3. Soil stiffness parameters in terms of soil friction angle and confining pressure

Parameter Relevant expression or value Reference


Sand relative density ðDr Þ f ½  Ibsen et al. (2009)
 1:8 s 30:28  1:53
15
 0:51
Sand modulus of elasticity ðEs Þ Otsmane and Amar Bouzid (2018)
2:93 Dr 1 þ 2k0
1;025 e s v0
3
Earth pressure coefficient at rest ðk0 Þ 1  sin f Jâky (1944)

Table 4. Parameters governing the hyperbolic model according to correlations and recommendations

Parameter Relevant expression or value Reference


Modulus exponent ðnÞ 0.51 Otsmane and Amar Bouzid (2018)
Failure ratio ðRf Þ 0.7 Wong and Duncan (1974)
 0:51
Unloading/reloading modulus number ðKur Þ 1 þ 2k0 Otsmane and Amar Bouzid (2018)
1;025 e2:93 Dr p0:49
a
3k0
Loading modulus number ðK Þ 0:667 Kur Duncan and Wong (1999)

© ASCE 04018141-10 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


ð ð
where pc ¼proper contribution of the slice itself; pr ¼contribution
Bt Dps B ai dv ¼ Nt bi dv þ pi (16)
v v of the preceding slice; fl ¼contribution of the subsequent slice; ai1 ,
ai , and aiþ1 are the element nodal displacement vectors of slices
where pi is the external forces vector to which slice i is subjected; i  1, i, and i þ 1, respectively; I = identity matrix; and Gi1 , Gi , and
and bi is the body forces vector that has the following compact form: Giþ1 are, respectively, the shear moduli at slices i  1, i, and i þ 1.
In Eq. (16) the matrix Dps is the matrix corresponding to a prob-
bi ¼ bpr pc fl
i  bi þ bi (17) lem of plane stresses, which may be given as
fl
2 3
with bpc pc pr
i ¼ L Nai ; bi ¼ L Nai1 ; bi ¼ L Naiþ1
pr fl
(18) 1 s 0
E 6 6 s 1 0 7 7
Dps ¼ 4 (21)
where Lpc ¼ lpc pr fl 1  s 1  s 5
i I; L ¼ li I and L ¼ li I
pr fl 2
(19) 0 0
2
2 Gi1 Gi 2 Gi Giþ1
and lpr
i ¼ ; lfl ¼ ;
ti ðGi1 ti þ Gi ti1 Þ i ti ðGi tiþ1 þ Giþ1 ti Þ From this fact, and unlike a fully 3D or plane strain problem, the
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

value of Poisson’s ratio equal to 0:5 is no longer a singular value.


and lpc pr fl
i ¼ li þ li (20) Eq. (17) shows that the fictitious body forces applied to a slice i
depend essentially on its own nodal displacements and on those of
slices sandwiching it. The numerical analysis of the vertical slicing
model has led to the familiar equations of a pseudo plane stress
Table 5. Reference wind turbine masses problem with body forces representing the interaction between the
slices forming the structure and its surrounding medium. The sub-
Turbine part Reference mass (t)
stitution of Eqs. (17) and (18) into Eq. (16) gives a more detailed
Rotor 110.0 governing equation
Nacelle 240.0 ð ð
Tower 347.5 ðBt D B þ Nt Lpc NÞ ai dv ¼ ðNt Lpr NÞai1 dv
v v
ð
Table 6. Reference wind turbine geometrical properties
þ ðNt Lfl NÞaiþ1 dv þ pi (22)
v

Tower property Reference (m)


In a more compact form Eq. (22) becomes
Tower height 87.6 fl
Tower diameter Si ai ¼ Hpr
i þ H i þ pi (23)
At the base 6.0
At the top 3.87
This equation cannot be solved as is because the right-hand
Tower wall thickness
terms are not available explicitly at the same time. Consequently, an
At the base 0.027
updating iterative process is needed
j j1
At the top 0.019 Sji aji ¼ Hpr fl
i þ Hi þ pi for j ¼ 1; 2; ………; jmax (24)

Fig. 9. Soil profile and monopile dimensions for the NREL-5 MW reference wind turbine. MSL = mean sea level.

© ASCE 04018141-11 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


where, j denotes the iteration number; and jmax = maximum number Eq. (25) is a hyperbola (Fig. 8), in which ðs 1  s 3 Þ is the princi-
of iterations allowed in the numerical process. pal stress difference; ðs 1  s 3 Þult is the asymptotic value of the
It is quite clear from Eq. (24) that the iterative nature is inherent principal stress difference at large axial strain; ɛ = axial strain; and
to the FE VSM even if the analysis is restricted to elastic deforma- Ei = initial tangent modulus.
tions. In this way, this equation is the backbone of the numerical On the basis of a wide variety of soil experiments performed by
procedure presented in this paper, which will be extended to encom- Janbu (1963), Duncan and Chang (1970) found that the initial modu-
pass the hyperbolic model next for analyzing nonlinear problems, lus Ei increases with the confining pressure s 3 . Consequently, they
particularly because the main purpose of this paper is to present the proposed for the initial tangent modulus the following formula:
monopiles in cohesionless soils.  n
Kondner (1963) and Kondner and Zelako (1963) approxi- s3
Ei ¼ KPa (26)
mated the relationship between the deviatoric stress and the axial Pa
strain for both clays and sands according to an analytic formula
as where K is a dimensionless factor termed “modulus number”; n is a
dimensionless parameter called “modulus exponent”; and Pa =
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

ɛ
ðs 1  s 3 Þ ¼ 1 ɛ
(25) atmospheric pressure (101:4 kPa) used to make K and n
þ nondimensional.
Ei ðs 1  s 3 Þult In the process of loading, the tangential stiffness of soil is often
required. This is given by
" #2  n
Table 7. Monopile properties Rf ð1  sin f Þðs 1  s 3 Þ s3
Et ¼ 1  KPa (27)
Monopile property Reference (m) 2c cos f þ 2s 3 sin f Pa
Monopile length
Embedment length 36.0 However, the value of Et cannot reflect the soil stiffness in
Overhang 30.0 cases of unloading/reloading; consequently, another expression is
Monopile diameter 6.0 needed. Assuming the same stress paths for both unloading and
Monopile wall thickness 0.06 reloading, Duncan and Chang (1970) proposed the following
expression:

25
H o r iz o n ta l fo r c e H ( M N )

W in k le r - R O W K S S ( S ø r e n s o n 2 0 1 2 )
W i n k le r - R O W K S S ( S ø r e n s o n e t a l. 2 0 1 0 )
W i n k l e r - R O W K S S ( W i e m a n n e t a l. 2 0 0 4 )
20 W in k le r -R O W K S S ( K a lle h a v e e t a l. 2 0 1 2 )

W in k le r -R O W K S S ( O 'N e ill a n d M u r c h is o n 1 9 8 3 )

W in k le r -R O W K S S ( R e e s e e t a l. 1 9 7 4 )
15 L P IL E ( Ju n g e t a l. 2 0 1 5 )

10

N A M P U L A L ( O ts m a n e a n d A m a r B o u z id 2 0 1 8 )
F E M ( Ju n g e t a l. 2 0 1 5 )

0
0 .0 0 0 .0 1 0 .0 2 0 .0 3 0 .0 4

M o n o p i l e h e a d d is p l a c e m e n t u ( m )

Fig. 10. Monopile head horizontal displacement u against H loading.

© ASCE 04018141-12 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


 n
s3 is generally sufficient to reach accurate solutions with acceptable
Eur ¼ Kur Pa (28)
Pa margins:
1. A number of 20 slices has been implemented in NAMPULAL.
where Eur = loading-unloading modulus; and Kur = corresponding This number, which has been set on the basis of a parametric
modulus number. In the Duncan-Chang basic model, the study involving many monopile behavior parameters (Amar
Poisson’s ratio s was assumed to be constant throughout the Bouzid et al. 2005a), has been found sufficient to model accu-
entire process. rately many soil–structure interaction problems (Amar Bouzid
To reduce the large number of experiment-based parameters et al. 2005b; Otsmane and Amar Bouzid 2018).
2. Unlike most implemented soil elastoplastic behavior criteria,
needed in the hyperbolic model, Otsmane and Amar Bouzid (2018)
which necessitate a significant number of iterations to subdue
established empirical relationships requiring only the soil friction
the out-of-balance equilibrium forces, the implemented hyper-
angle and the confining pressure s 3 . These relationships are listed
bolic model in NAMPULAL requires only two iterations. This
in Tables 3 and 4. [For more details the reader is referred to
fact alleviates considerably the entire process of the solution
Otsmane and Amar Bouzid (2018)].
and makes it easier to find fast solutions, even for the most
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

These expressions have been implemented in the FE com-


complex soil/structure interaction problems.
puter code NAMPULAL, which will be described in the next 3. In addition to the rectangular cross-sectional monopiles that are
subsection. automatically taken into account due to the shape of the vertical
slice, the solid circular or tubular cross-section monopiles are
easily dealt with by prescribing the effective bending stiffness.
Computer Program NAMPULAL Used in This Work Hence, an equivalent Young’s modulus is extracted according
A Fortran computer program called NAMPULAL for the analysis to the following formula:
of axially and laterally loaded single monopiles has been written
on the basis of the FE vertical slices equations described earlier 192 ðEI Þact
Epeq ¼ (29)
[Eqs. (14)–(24)] in conjunction with those of hyperbolic model p 2 ðDp Þ4
equations. Only the features of this computer program are given
here. Although the computational process in NAMPULAL is nat- This expression has been set on the assumption that the square
urally iterative to fulfill the slices equilibrium, it does not require cross-sectional monopile under consideration in NAMPULAL
a significant number of iterations to reach convergence. For the has the same cross-sectional area as the effective circular cross-
problems analyzed so far, a number revolving around 20 iterations sectional monopile.

25
H o r iz o n t a l f o r c e H ( M N )

W i n k le r - R O W K S S ( S ø r e n s o n 2 0 1 2 )
W in k le r - R O W K S S ( S ø r e n s o n e t a l. 2 0 1 0 )

W i n k l e r - R O W K S S ( W i e m a n n e t a l. 2 0 0 4 )
20
W in k le r - R O W K S S ( K a lle h a v e e t a l. 2 0 1 2 )

W in k le r -R O W K S S ( O 'N e ill a n d M u r c h is o n 1 9 8 3 )

W in k le r -R O W K S S ( R e e s e e t a l. 1 9 7 4 )
15
L P IL E ( Ju n g e t a l. 2 0 1 5 )

10

N A M P U L A L ( O ts m a n e a n d A m a r B o u z id 2 0 1 8 )

F E M ( Ju n g e t a l. 2 0 1 5 )

0
0 .0 0 0 .0 4 0 .0 8 0 .1 2 0 .1 6

M o n o p ile h e a d r o t a tio n • ( °)

Fig. 11. Monopile head rotation u against H loading.

© ASCE 04018141-13 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


NAMPULAL for FE Modeling against The flexibility factors are found first and then inversed to obtain
Winkler-ROWKSS for BNWF Analysis the stiffness coefficients of Eq. (31).

Through the analysis of two case histories the monopile head Verifying 5-MW Reference Wind Turbine from National
stiffnesses provided by the FEM using NAMPULAL are com- Renewable Energy Laboratory
pared in this section with those given by the BNWF model using
the computer code Winkler-ROWKSS. The monopile head move- To assess NAMPULAL’s FE results compared with those of the
ments (lateral displacements and rotations) are the key elements API and the recently developed p-y curves implemented in
for the determination of the monopile head stiffness, which can Winkler-ROWKSS, it is useful to consider the National Renewable
be quantified by three springs, two for controlling horizontal and Energy Laboratory (NREL) 5-MW baseline wind turbine, which is
rocking movements and one for the cross-coupling interaction. mounted atop a monopile foundation at a 20-m water depth
The relationship between lateral stiffness KL , rocking stiffness (Jonkman et al. 2009; Jung et al. 2015).
KR , and cross-coupling stiffness KLR may be expressed in a matrix To aid concept studies and research activities on offshore
wind energy, the NREL developed a reference wind turbine that
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

form as
closely represents utility-scale OWTs found in the today’s mar-

 ket. The OWT properties have been obtained from publicly avail-
H KL KLR uL
¼ (30) able information on real wind turbine prototypes and conceptual
M KRL KR uR
models. The NREL 5-MW mass and structural details are listed,
respectively, in Tables 5 and 6. Fig. 9 shows a schematic of the
where H and M = shear force and the overturning moment applied reference wind turbine. It has three blades with variable speed
at the monopile head, respectively; and uL and u R = lateral displace- and pitch control. The rotor diameter is 126 m. The support struc-
ment and rotation of the monopile head, respectively. tures used for OWTs include monopiles, gravity bases, space
The stiffness coefficients appearing in Eq. (30) are related to frames, and floating structures. In this study, for a monopile foun-
their inverse flexibility coefficients by the following expres- dation investigation, a representative soil profile was assumed,
sions: which was adopted from the work of Passon (2006), as shown in
Fig. 9. This profile consists of three sand layers with the same unit
IR IL ILR
KL ¼ KR ¼ ; KLR ¼ (31) weight and three different friction angles. The geometrical char-
IL IR  ILR
2 IL IR  ILR
2 IL IR  ILR
2
acteristics are given in Table 7.

200
A p p l i e d o v e r t u r n in g m o m e n t M ( M N .m )

W i n k le r - R O W K S S ( S ø r e n s o n 2 0 1 2 )

W in k le r - R O W K S S ( S ø r e n s o n e t a l. 2 0 1 0 )

W i n k le r - R O W K S S ( W i e m a n n e t a l. 2 0 0 4 )
W i n k le r - R O W K S S ( K a ll e h a v e e t a l. 2 0 1 2 )
150
W in k le r - R O W K S S ( R e e s e e t a l . 1 9 7 4 )

W in k le r - R O W K S S ( O 'N e i l l a n d M u r c h i s o n 1 9 8 3 )

L P IL E ( Ju n g e t a l. 2 0 1 5 )

100

50

N A M P U L A L ( O ts m a n e a n d A m a r B o u z id 2 0 1 8 )

F E M ( J u n g e t a l. 2 0 1 5 )

0
0 .0 0 0 .0 4 0 .0 8 0 .1 2 0 .1 6
M o n o p il e h e a d r o t a t i o n • ( ° )

Fig. 12. Monopile head rotation u against M loading.

© ASCE 04018141-14 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


The numerical investigation of this reference turbine was per-

0.0000135
0.0000943
et al. (2012)
Kallehave
formed by examining the monopile head movement (displacements

0.0011

2.233

−15.591
182.897
and rotations) evolutions with the increasing applied loading at the
monopile top, as indicated in Figs. 10–12. Fig. 10 shows the varia-
tion of u with H, whereas Fig. 11 is dedicated to the evolution of u
with H and Fig. 12 to u with M. The initial stiffnesses of these
curves have been used for the determination for monopile head flex-

O’Neill and Murchison


ibility coefficients first, and then stiffness coefficients that appear in

0.0000154
Table 8. The latter are the key elements for some dynamic charac-

0.00013
(1983)

0.0017
teristics, such as OWT natural frequency, if the soil–monopile inter-

−13.232
174.218
1.604
action is considered.
At first it is easy to notice that the curves are distinguished
into four groups. In the first group, which gathers the FE results,
a perfect agreement has been obtained between NAMPULAL’s
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

results and those of Jung et al. (2015) who used the commercial
FE package ABAQUS in their computations. This means that the

0.0000154
Reese et al.
computer code NAMPULAL, which has been validated against

0.00013
(1974)

0.0017

−13.232
174.218
1.604
different FE packages (Otsmane and Amar Bouzid 2018), once
again proved its validity in dealing with large-diameter monop-
iles in sandy deposits. In the second group, which is formed by
Winkler models adopted by the API method and based on the old-

0.0000187
est versions of p-y curves (Reese et al. 1974; O’Neill and

0.00317

0.00019
Sørensen
(2012)
Murchison 1983) (LPILE), the results are nearly identical. This is

134.295
0.793

−8.000
obvious because the three models were originally based on the
same work by Reese et al. (1974). These models significantly
underestimate the monopile head lateral displacement and show
that they are not suitable for large-diameter monopiles. This

0.0000146
LPILE (Jung
et al. 2015)

0.00012
underestimation leads to an overestimation in soil–monopile

0.0016

193.819
1.816

−15.092
stiffness, which wrongfully affects the dynamic properties of the
OWT. The third group is relevant to Winkler models, which used
the latest propositions to enhance p-y curves. In this group it is
easy to see that results provided by Winkler-ROWKSS using
Sørensen et al. (2010), Sørensen (2012), and Wiemann et al.

0.0000177
FEM (Jung
et al. 2015)

0.00019
(2004) curves are in close agreement with each other, although

0.0037

123.211
0.583

−6.236
they have been formulated differently, and are situated halfway
between the API and the FE models. It seems that the p-y modifi-
cations have brought some improvement toward a more accurate
monopile head displacement. Finally, in the last group, which
Wiemann et al.

consists of the Kallehave et al. (2012) model alone, the results 0.0000174
are heavily inaccurate. Lateral displacements are significantly 0.00016
(2004)

0.0024

1.086

−10.055
150.458
underestimated, which renders the method unsatisfactory to
design laterally loaded monopiles. The model by Kallehave et al.
(2012), presumably, has been calibrated only for the monopiles
they tested and on which the p-y expression was formulated, but
the model had not been verified for a large database of field
Sørensen et al.

0.0000176
0.00016
(2010)

experiments.
0.0026

0.949

−8.946
141.360

Analysis of Monopile Response at Horns Rev (Denmark)


Horns Rev is a site located in the Danish sector of the North Sea, in
Table 8. Flexibility and stiffness coefficients

which 80 Vestas V80 OWTs are installed. The turbines of this wind
NAMPULAL

0.0000187
0.00360

0.00019

farm are founded on steel monopiles with 4-m outer diameters and
117.316
0.608

−6.225

lengths varying between 30 and 32:7 m. The transition from the


tower to the monopile is secured by a transition piece whose outer
and inner diameters are, respectively, 4:24 and 4:15 m. The
monopile, subject of the present work, has an outer diameter of
4 m, a length of 31:6 m, and a variable wall thickness wt. Hence,
Flexibility coefficients

the flexural stiffness ðEI Þp varies along the monopile, as shown in


Stiffness coefficients

KR ðGN=m=radÞ

Fig. 13. The monopile has been driven to 31:8 m below the mean
IR ðrad=GN=mÞ

sea level (MSL), which results in an embedded depth equal to


KL ðGN=mÞ
ILR ð1=GNÞ
IL ðm=GNÞ

KLR ðGNÞ

21:9 m.
Coefficient

The soil profile at the location of the considered monopile con-


sists primarily of sand and it can roughly be divided into six differ-
ent layers. The first layer 6:5 m below the seabed is sand, followed

© ASCE 04018141-15 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


- 0.2 m
MSL 0.0

11.8
- 9.9 m

= 50 ,
4.5
( ) = 2.54 108 2

7.8 2.0
= 52 ,
( ) = 2.64 108 2
5.4

= 50 , 4.0
2.1
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

8 2
( ) = 2.54 10

= 40 , 4.0 / / 4.2
( ) = 2.05 108 2
4.0
3.7
= 30 4.0 - 31.8 m
( ) = 1.55 108 2

Fig. 13. Details of the monopile at Horns Rev.

Table 9. Soil strata including average values of the strength and stiffness parameters for each soil layer at Horns Rev
 
Soil layer Type Depth (m) Es ðMPaÞ g ð g Þ kN=m3 f ðdegreesÞ c ðdegreesÞ s
1 Sand 0.0–4.5 130.0 20(10) 45.4 15.4 0.28
2 Sand 4.5–6.5 114.3 20(10) 40.7 10.7 0.28
3 Sand to silty sand 6.5–11.9 100.0 20(10) 38.0 8.0 0.28
4 Sand to silty sand 11.9–14.0 104.5 20(10) 36.6 6.6 0.28
5 Sand/silt/organic 14.0–18.2 4.5 17(7) 27.0 0.0 0.28
6 Sand > 18:2 168.8 20(10) 38.7 8.7 0.28

0 .0 6
M o n o p ile d is p la c e m e n t ( m )

O 'N e i ll a n d M u r c h i s o n ( 1 9 8 3 )
K a ll e h a v e e t a l . ( 2 0 1 2 )
0 .0 5 3D
F la c ( A u g u s t e s e n e t a l. 2 0 0 9 )

0 .0 4 W ie m a n n e t a l. ( 2 0 0 4 )
N A M P U L A L ( O t s m a n e a n d A m a r B o u z id 2 0 1 8 )
0 .0 3 S o r e n s o n e t a l. 2 0 1 0

Soren son 2 012


0 .0 2
R e e s e e t a l. 1 9 7 4

0 .0 1

0 .0 0

-0 . 0 1
P o in t s o f z e r o d is p la c e m e n t s

-0 . 0 2
0 .0 2 .5 5 .0 7 .5 1 0 .0 1 2 .5 1 5 .0 1 7 .5 2 0 .0 2 2 .5

D ep th (m )

Fig. 14. Monopile deflected shapes according to different models at Horns Rev.

© ASCE 04018141-16 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


by 7:5 m of silty sand. From 14 to 18:2 m below the seabed, silt/ observed at the monopile top and tip. This is probably because of
sand including organic material dominates the soil profile. Under the soil failure criteria adopted in both models. Here I believe that
this layer the sand is found. The soil conditions are summarized in the FE analysis is more accurate because the monopile behaves as
Table 9. a stiff monopile with more pronounced displacements at the
The monopile is subjected to a static horizontal load H ¼ monopile tip.
4:6 MN and a bending moment M ¼ 95 MN=m, both acting at Further description of Fig. 14 can be made in three important
seabed level. In this study, the monopile displacements profile points. First, Winkler models using either Reese et al. (1974) or
has been chosen as a comparison parameter, on which the FE O’Neill and Murchison (1983) p-y curves significantly underesti-
models are set against those of Winkler-ROWKSS. In addition to mate the displacement profile along the entire monopile length
results of these models, results provided by the commercial pack- producing only small values of displacements at the monopile
age FLAC3D have been added to the graphs to assess the accu- tip. This means that the behavior in these models tends to mimic
racy of FE analysis by NAMPULAL. Numerical computations that of slender monopiles. Furthermore, the point of zero dis-
using FLAC3D have been performed by Augustesen et al. (2009) placement is located approximately at midlength, which is incon-
who considered the sand as an elastoplastic material obeying the sistent with the shape pattern of a stiff monopile. Second, and
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

Mohr-Coulomb failure criterion. Results of the comparison are except the Winkler model based on the Kallehave et al. (2012)
shown in Fig. 14. p-y curve, the Winkler models using the other enhanced p-y
At first glance, NAMPULAL’s displacements are almost iden- curves tend to provide more accurate results at the upper part of
tical to those of FLAC3D, not only for the point of zero displace- the monopile. Results are located halfway between those of the
ment but for the entire profile. However, slight differences are API and FEM; however, they show inconsistencies with the
deformed shapes of a stiff monopile behavior. The most accu-
rate model seems to be the method based on the Sørensen et al.
(2010) p-y curve because it perfectly matches the FE model at
the monopile head, but it significantly overestimates the lateral
displacements at the monopile tip. Third, it is quite clear that
displacements by the Winkler model based on the p-y curve for-
mulated by Kallehave et al. (2012) are importantly incorrect
because they are much lower than even those of the API models.
At the monopile head the displacement using this model is 2.5
times less than that of NAMPULAL. This confirms again that
this method is not appropriate for designing large-diameter
monopiles.

Conclusions

For decades the BNWF model has been a successful tool for design-
Fig. 15. Soil stiffness variation with depth according to three patterns. ing piles supporting offshore platforms in the oil and gas industry
and any other sector of civil engineering, such as tall buildings or

Direction of
movement

Shear failure surface


Movement of block

Direction of
movement
(a) (b)

Fig. 16. Soil failure models proposed by Reese et al. (1974): (a) shallow wedge model; and (b) deep horizontal flow model. (Data from Reese et al.
1974.)

© ASCE 04018141-17 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


bridges. The incapability of this model to design large-diameter eral displacement profile was studied. The major conclusions
monopiles supporting OWTs has been practically proven for the reached were as follows:
overwhelming number of inconsistencies that characterize the back- 1. The monopile head displacements in the FE analysis provided
ground theories on which its p-y curves have been formulated on by NAMPULAL are reasonably accurate. This means that the
one hand and for the limitations inherent to the BNWF model itself monopile head stiffnesses KLRigid ; KRRigid , and KLR
Rigid
, which are
on the other hand. required to compute some dynamic characteristics of OWTs,
After a short description of how the failure mechanism of a are sufficiently precise, and the soil–monopile interaction is
small-diameter monopile differs from that of a large-diameter properly considered. Furthermore, the displacement profile in
monopile, the p-y curves developed in the late 1970s and early this numerical method is consistent with the deformed shape of
1980s (adopted by the API) have been thoroughly criticized, and a stiff monopile.
their weak points as well as those found in the BNWF model itself 2. The BNWF models based on the Reese et al. (1974) or the
were described in three subsections. Modifications to enhance the O’Neill and Murchison (1983) p-y curves or even LPILE,
performance of the p-y curves proposed recently were also given. which is based on these curves, significantly overestimate the
The targets in this paper were to confirm the convenience of monopile head stiffness. Consequently, they are not able to
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

the FEM in analyzing large-diameter monopiles under horizontal quantify correctly the soil–monopile interaction.
loading and how inappropriate the BNWF model is when dealing 3. Regarding monopile head movements, the BNWF models
with the previously mentioned subject. These objectives were based on Wiemann et al. (2004), Sørensen et al. (2010), and
achieved by considering two computer programs. The first pro- Sørensen (2012) p-y curves can be considered moderately pre-
gram is a FE computer code called NAMPULAL, which has been cise because they are close to the FE outcomes. They can be
detailed in Otsmane and Amar Bouzid 2018. This code is based used to quantify the monopile head stiffnesses with an accepta-
on a special treatment of soil–structure interaction problems and ble margin. However, the monopile deformed shapes in these
uses the hyperbolic model as a soil failure criterion. The second models are not consistent with large-diameter rigid monopiles.
program, which bears the acronym Winkler-ROWKSS, was spe- 4. The model based on the Kallehave et al. (2012) p-y curve is not
cifically written for the purpose of this paper by extending an appropriate for dealing with large-diameter monopiles.
existing Fortran computer code to encompass the API curves as Although significant progress has been made to enhance the
well as those recently proposed to improve the poor performance poor performance of BNWFs by suggesting new p-y curves in
of the BNWF approach. Two case histories were studied: the first which very careful experiments or powerful numerical tools have
is a monopile embedded in a multilayer sandy soil, which serves been performed, the behavior of the monopiles has not been mas-
as a support foundation to the 5-MW NREL wind turbine, and the tered and inconsistencies in their deformations still exist. This, in
second is a monopile supporting OWTs at Horns Rev in the the author’s opinion, is not due to the inaccuracies in the proposed
Danish sector of the North Sea. At the first site, the evolution of p-y curves but to the model itself, which ignores certain soil–
the monopile head movements with applied loading provided by monopile interaction aspects, such as the vertical shear stresses
the different models of BNWF as well as the FEM were examined induced on the external monopile perimeter and the moment and
and compared. However, at the second site, only the monopile lat- horizontal force developed across the monopile base.

Appendix I. Impedance Functions for Flexible and Rigid Monopiles Embedded in Three Soil Profiles
The authors whose expressions appear in these tables considered three soil profiles in which the soil stiffness is constant with depth, varies
linearly, or varies as a square root of depth. EsD is Young’s modulus of the soil at one pile diameter (Dp Þ depth at which the three patterns of
variation meet (Fig. 15). The following table includes the impedance functions for slender piles proposed by different authors in three soil
profiles. Most expressions were deduced from flexibility coefficients computed by the different authors in their papers.

Authors KLFlexible =EsD Dp KLR


Flexible
=EsD D2p KRFlexible =EsD D3p
Homogeneous soil
 0:143  0:428  0:714
Randolph (1981) 1:630 E =E 0:342 E =E 0:197 E =E
 p sD 0:213  p sD 0:479  p sD 0:695
Amar Bouzid (1997) 1:152 E =E 0:317 E =E 0:294 E =E
 p sD 0:210  p sD 0:500  p sD 0:750
Gazetas (1991) 1:000 Ep =EsD 0:220 E =E 0:150 E =E
 0:182  p sD 0:454  p sD 0:727
Davies and Budhu (1986) 1:292 E =E 0:309 E =E 0:182 E =E
 p sD 0:186  p sD 0:500  p sD 0:730
Shadlou and Bhattacharya (2016) 1:580 Ep =EsD 0:327 Ep =EsD 0:204 Ep =EsD
Gibson’s soil
 0:333  0:555  0:777
Randolph (1981) 0:853 Ep =EsD 0:313 E =E 0:195 Ep =EsD
 0:333  p sD 0:555  0:777
Budhu and Davies (1987) 0:734 E =E 0:270 E =E 0:173 Ep =EsD
 p sD 0:327  p sD 0:523  0:726
Amar Bouzid (1997) 1:138 E =E 0:591 E =E 0:450 Ep =EsD
 p sD 0:350  p sD 0:600  0:800
Gazetas (1991) 0:600 Ep =EsD 0:170 E =E 0:150 Ep =EsD
 0:340  p sD 0:567  0:780
Shadlou and Bhattacharya (2016) 0:861 Ep =EsD 0:291 Ep =EsD 0:188 Ep =EsD
Parabolic nonhomogeneity
 0:280  0:530  0:770
Gazetas (1991) 0:800 Ep =EsD 0:240 Ep =EsD 0:150 Ep =EsD
 0:270  0:520  0:760
Shadlou and Bhattacharya (2016) 1:106 Ep =EsD 0:319 Ep =EsD 0:193 Ep =EsD

However, expressions in the following table, which includes previous research impedance functions for stiff monopiles proposed by dif-
ferent authors in three soil profiles, are listed as they appeared in their respective publications:

© ASCE 04018141-18 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


Authors KLRigid =EsD Dp Rigid
KLR =EsD D2p KRRigid =EsD D3p
Homogeneous soil
 0:627  1:483  2:049
Carter and Kulhawy (1992) 1:884 Lp =Dp 1:048 Lp =Dp 1:910 Lp =Dp
 0:710  1:670  2:459
Higgins et al. (2013) 2:426 Lp =Dp 1:440 Lp =Dp 1:789 Lp =Dp
 0:620  1:560  2:500
Shadlou and Bhattacharya (2016) 3:488 Lp =Dp 1:935 Lp =Dp 1:798 Lp =Dp
 0:668  1:636  2:495
Aissa et al. (2017) 2:756 Lp =Dp 1:595 Lp =Dp 1:731 Lp =Dp
Gibson’s soil
 2:041  3:061  3:941
Higgins et al. (2013) 0:929 Lp =Dp 0:633 Lp =Dp 0:672 Lp =Dp
 1:661  2:665  3:605
Abed et al. (2016) 1:708 Lp =Dp 1:233 Lp =Dp 1:153 Lp =Dp
 1:530  2:500  3:450
Shadlou and Bhattacharya (2016) 2:561 Lp =Dp 1:935 Lp =Dp 1:730 Lp =Dp
Soil with parabolic variation of stiffness
 0:997  1:767  2:562
Abed et al. (2016) 2:841 Lp =Dp 2:933 Lp =Dp 3:894 Lp =Dp
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

 1:070  2:000  3: 000


Shadlou and Bhattacharya (2016) 2:905 Lp =Dp 1:962 Lp =Dp 1:771 Lp =Dp

Appendix II. Analytical Expression for the Depth   1 þ tan f


Separating the Two Failure Models as Proposed by p f 2;
c2 ¼ tan þ ¼
Reese et al. (1974) 4 2 f
1  tan
2
Soil deformation models proposed by Reese et al. (1974) are illus-
 2
trated by Fig. 16. f f
The ultimate resistance at a shallow depth [Fig. 16(a)] is given 1 þ tan tan
f 2 c1 c5 2 2 ;
by c1 c2 ¼ 2tan 
3
4 ; ¼ pffiffiffi 
2 f c6 2 f

1  tan 1  tan
k0 ZWF tan f sin b tan b 2 2
Pct ¼ g ZWF þ
tanð b  f Þcosa tanð b  f Þ !2
f 1 þ tan f2
c2 c3 ¼ tan
2
; and
ðDp þ ZWF tan b tanaÞ þ k0 ZWF tan b ðtan f sin b  tanaÞ  ka Dp 2 1  tan f2
0 1
(32) f
fB2 cos
2  1C
c1 c5  c3 ¼ tan B @ pffiffiffi C
A
Whereas in depth the ultimate soil resistance is reached by a 2 2 1  tan f
flow around the pile [Fig. 16(b)] 2
 
Pcd ¼ ka Dp g ZWF tan8 b  1 þ k0 Dp g ZWF tan f tan4 b (33) These formulas allow Eq. (35) to be written as

where k0 = earth pressure coefficient at rest; Dp = pile diameter; a5 a2 a


and f = internal friction angle of sand. þ 2k0 4 
b5 ð1  bÞ b bð1  bÞ
Because a ¼ f =2 and b ¼ p =4 þ f =2, it is useful to use ZWF ¼ Dp   (36)
2k0 a2 2d
other simplifying expressions: c1 ¼ tan f , c2 ¼ tan b ¼ tanðp = pffiffiffi þ 2 þ k0 pffiffiffi  1
4 þ f =2Þ, c3 ¼ tana ¼ tanð f =2Þ, c4 ¼ tanð b  f Þ ¼ tanðp = 2b b 2b
4  f =2Þ; c5 ¼ sin b ¼ sinðp =4 þ f =2Þ; and c6 ¼ cosa ¼ cos
f =2. The active pressure coefficient is ka ¼ tan2 ðp =4  f =2Þ ¼ c24 . where a ¼ 1 þ tanð f =2Þ; b ¼ 1  tanð f =2Þ; and d ¼ cos
After writing Eqs. (32) and (33) in terms of c1  c6 , and equat- ð f =2Þ.
ing these expressions, the depth ZWF , at which the wedge failure
model ends, will have the following expression: References

c24 c72  c12 þ k0 c1 c32 þ c34  c2 Abdel-Rahman, K., and M. Achmus, 2005. “Finite element modeling of
ZWF ¼ Dp c1 c5 (34) horizontally loaded monopile foundations for offshore wind energy con-
k0 þ c22 c3 þ k0 ðc1 c5  c3 Þ verters in Germany.” In Proc., Int. Symp. on Frontiers in Offshore
c6 Geotechnics (ISFOG), 391–396. Perth, Australia: CRC Press.
Abed, Y., D. Amar Bouzid, S. Bhattacharya, and M. H. Aissa. 2016. “Static
Since the coefficient c4 is the reciprocal of c2 , then c2 c4 ¼ 1. impedance functions for monopiles supporting offshore wind turbines
Eq. (34) becomes in nonhomogeneous soils-emphasis on soil/monopile interface charac-
teristics.” Earthquakes Struct. 10 (5): 1143–1179. https://doi.org/10
c52 þ k0 c1 c32  c2 .12989/eas.2016.10.5.1143.
ZWF ¼ Dp c1 c5 (35) Achmus, M., K.-S. Kuo, and K. Abdel-Rahman. 2009. “Behavior of
k0 þ c22 c3 þ k0 ðc1 c5  c3 Þ monopile foundations under cyclic lateral load.” Comput. Geotech. 36
c6
(5): 725–735. https://doi.org/10.1016/j.compgeo.2008.12.003.
Adhikari, S., and S. Bhattacharya. 2011. “Vibrations of wind-turbines con-
To use only the angle of friction f , the constant groups in Eq. sidering soil-structure interaction.” Wind Struct. 14 (2): 85–112. https://
(35) should be written in function of f alone doi.org/10.12989/was.2011.14.2.085.

© ASCE 04018141-19 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


Adhikari, S., and S. Bhattacharya. 2012. “Dynamic analysis of wind turbine DNV (Det Norske Veritas). 2013. Offshore standard: Design of offshore
towers on flexible foundations.” Shock Vib. 19 (1): 37–56. https://doi wind turbine structures. DNV-OS-J101. Høvik, Norway: DNV.
.org/10.1155/2012/408493. Duncan, J., and C. Chang. 1970. “Nonlinear analysis of stress and strain in
Ahmed, S., and B. Hawlader. 2016. “Numerical analysis of large-diameter soils.” J. Soil Mech. Found. Div. 96 (SM5): 1629–1653.
monopiles in dense sand supporting offshore wind turbines.” Int. J. Duncan, J. M., and K. S. Wong. 1999. User’s manual for SAGE. Vol. II—
Geomech. 16 (5): 04016018. https://doi.org/10.1061/(ASCE)GM.1943 Soil properties manual. Blacksburg, VA: Centre for Geotechnical
-5622.0000633. Practice and Research, Virginia Tech.
Aissa, M. H., D. Amar Bouzid, and S. Bhattacharya. 2017. “Monopile Fan, C.-C., and J. H. Long. 2005. “Assessment of existing methods for pre-
head stiffness for serviceability limit state calculations in assessing dicting soil response of laterally loaded piles in sand.” Comput.
the natural frequency of offshore wind turbines.” Int. J. Geotech. Eng. Geotech. 32 (4): 274–289. https://doi.org/10.1016/j.compgeo.2005.02
12 (3): 267–283. .004.
Amar Bouzid, D. 1997. “Analyse semi-analytique par elements finis des Galvín, P., A. Romero, M. Solís, and J. Domínguez. 2017. “Dynamic char-
pieux isoles sollicites horizontalement dans un milieu a comporte- acterisation of wind turbine towers account for a monopile foundation
ment non-lineaire.” M.S. thesis, École Nationale Polytechnique. and different soil conditions.” Struct. Infrastruct. Eng. 13 (7): 942–954.
Amar Bouzid, D., B. Tiliouine, and P. A. Vermeer. 2004. “Exact formula- https://doi.org/10.1080/15732479.2016.1227342.
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

tion of interface stiffness matrix for axisymmetric bodies under non-axi- Gazetas, G. 1991. “Foundation vibrations.” In Foundation engineering
symmetric loading.” Comput. Geotech. 31 (2): 75–87. https://doi.org handbook. 2nd ed. Edited by H.-Y. Fang, 553–593. New York: Van
/10.1016/j.compgeo.2004.01.007. Nostrand Reinhold.
Amar Bouzid, D., P. A. Vermeer, and B. Tiliouine. 2005a. “Finite element Hald, T., C. Mørch, L. Jensen, C. L. Bakmar, and K. Ahle, 2009.
vertical slices model: Validation and application to an embedded square “Revisiting monopile design using p-y curves–results from full scale
footing under combined loading.” Comput. Geotech. 32 (2): 72–91. measurements on Horns Rev.” In Proc., of the European Offshore Wind
https://doi.org/10.1016/j.compgeo.2005.01.001. Conf. Stockholm, Sweden: WindEurope.
Amar Bouzid, D., P. A. Vermeer, B. Tiliouine, and M. Mir. 2005b. “An effi- Higgins, W., C. Vasquez, D. Basu, and D. Griffiths. 2013. “Elastic solutions
cient pseudo three-dimensional finite element model: Presentation and for laterally loaded piles.” J. Geotech. Geoenviron. Eng. 139 (7): 1096–
analysis of two soil/foundation interaction problems.” Int. J. Comput. 2 1103. https://doi.org/10.1061/(ASCE)GT.1943-5606.0000828.
(2): 231–253. https://doi.org/10.1142/S0219876205000454. Ibsen, L. B., M. Hansen, T. H. Hjort, and M. Thaarup, 2009. MC-parameter
API (American Petroleum Institute). 2014. Geotechnical and foundation calibration of Baskarp sand No. 15. DCE Technical Rep. No. 62.
design considerations. API RP 2GEO. Washington, DC: API. Aalborg, Denmark: Dept. of Civil Engineering, Univ. of Aalborg.
Arany, L., S. Bhattacharya, J. H. G. Macdonald, and S. J. Hogan. 2016. Jâky, J. A. 1944. “Nyugalmi nyomâs tenyezöje” [The coefficient of earth
“Closed form solution of Eigen frequency of monopile supported off- pressure at rest].” J. Soc. Hungarian Archit. Eng. 355–358.
Janbu, N. 1963. “Soil compressibility as determined by oedometer and tri-
shore wind turbines in deeper waters incorporating stiffness of substruc-
axial tests.” In Vol. 1 of Proc., European Conf. on Soil Mechanics and
ture and SSI.” Soil Dyn. Earthquake Eng. 83 (Apr): 18–32. https://doi
Foundation Engineering, 19–25. Wiesbaden, Germany: Deutsche
.org/10.1016/j.soildyn.2015.12.011.
Gesellschaft für Erd-und Grundbau.
Arany, L., S. Bhattacharya, J. H. G. Macdonald, and S. J. Hogan. 2017.
Jonkman, J., S. Butterfield, W. Musial, and J. Scott, 2009. Reference wind
“Design of monopiles for offshore wind turbines in 10 steps.” Soil Dyn.
turbine for offshore system development. Technical Rep. No. NREL/TP-
Earthquake Eng. 92 (Jan): 126–152. https://doi.org/10.1016/j.soildyn
500-38060. Golden, CO: National Renewable Energy Laboratory.
.2016.09.024.
Jung, S., S. R. Kim, A. Patil, and L. C. Hung. 2015. “Effect of monopile
Ashford, S., and T. Juirnarongrit. 2003. “Evaluation of pile diameter effect
foundation modeling on the structural response of a 5-MW wind turbine
on initial modulus of subgrade reaction.” J. Geotech. Geoenviron. Eng.
tower.” Ocean Eng. 109 (Nov): 479–488. https://doi.org/10.1016/j
129 (3): 234–242. https://doi.org/10.1061/(ASCE)1090-0241(2003)129: .oceaneng.2015.09.033.
3(234). Kallehave, D., C. T. LeBlanc, and M. A. Liingaard, 2012. “Modification of
Ashour, M., and G. Norris. 2000. “Modeling lateral soil-pile response based the API p-y formulation of initial stiffness of sand.” In Proc., 7th Int.
on soil-pile interaction.” J. Geotech. Geoenviron. Eng. 126 (5): 420– Conf. on Offshore Site Investigations and Geotechnics, 465–472. Royal
428. https://doi.org/10.1061/(ASCE)1090-0241(2000)126:5(420). Geographical Society, London: Society for Underwater Technology.
Augustesen, A. H., K. T. Brodbaek, M. Moller, S. P. H. Sorensen, L. B. Kondner, R. 1963. “Hyperbolic stress-strain response: Cohesive soils.” J.
Ibsen, T. S. Pedersen, and L. Andersen, 2009. “Numerical modelling of Soil Mech. Found. Div. 89 (SM1): 115–143.
large-diameter steel piles at Horns Rev.” In Proc., Twelfth Int. Conf. on Kondner, R. L., and J. S. Zelako, 1963. “A hyperbolic stress-strain formula-
Civil, Structural and Environmental Engineering Computing, edited by tion for sands.” In Vol. 1 of Proc., 2nd Pan American Conf. on Soil
B. H. V. Topping, L. F. Costa Neves, and R. C. Barros. Stirlingshire, Mechanics and Foundation Engineering, 289–324. Sao Paulo, Brazil:
Scotland: Civil-Comp Press. Brazilian Association of Soil Mechanics.
Bisoi, S., and S. Haldar. 2015. “Design of monopile supported offshore Lombardi, D., S. Bhattacharya, and D. M. Wood. 2013. “Dynamic soil–
wind turbine in clay considering dynamic soil–structure-interaction.” structure interaction of monopile supported wind turbines in cohesive
Soil Dyn. Earthquake Eng. 73 (June): 103–117. https://doi.org/10.1016 soil.” Soil Dyn. Earthquake Eng. 49 (June): 165–180. https://doi.org/10
/j.soildyn.2015.02.017. .1016/j.soildyn.2013.01.015.
Budhu, M., and T. G. Davies. 1987. “Nonlinear analysis of laterality loaded Mayne, P., and F. Kulhawy. 1982. “K0-OCR relationships in soil.” J.
piles in cohesionless soils.” Can. Geotech. J. 24 (2): 289–296. https:// Geotech. Engrg. Div. 108 (GT6): 851–872.
doi.org/10.1139/t87-034. Michalowski, R. 2005. “Coefficient of earth pressure at rest.” J. Geotech.
Carswell, W., S. R. Arwade, D. J. DeGroot, and M. A. Lackner. 2015. Geoenviron. Eng. 131 (11): 1429–1433. https://doi.org/10.1061
“Soil-structure reliability of offshore wind turbine monopile founda- /(ASCE)1090-0241(2005)131:11(1429).
tions.” Wind Energy 18 (3): 483–498. https://doi.org/10.1002/we.1710. O’Neill, M. W., and J. M. Murchison, 1983. An evaluation of p-y relation-
Carter, J., and F. Kulhawy. 1992. “Analysis of laterally loaded shafts in ships in sands. Research Rep. No. GT-DF02-83. Houston: Dept. of
rock.” J. Geotech. Eng. 118 (6): 839–855. https://doi.org/10.1061 Civil Engineering, Univ. of Houston.
/(ASCE)0733-9410(1992)118:6(839). Otsmane, L., and D. Amar Bouzid. 2018. “An efficient FE model for SSI:
Davies, T. G., and M. Budhu. 1986. “Non-linear analysis of laterally loaded Theoretical background and assessment by predicting the response of
piles in heavily overconsolidated clays.” Geotechnique 36 (4): 527–538. large diameter monopiles supporting OWECs.” Comput. Geotech. 97
https://doi.org/10.1680/geot.1986.36.4.527. (May): 155–166. https://doi.org/10.1016/j.compgeo.2017.12.001.
Dickin, E. A., and M. Laman. 2003. “Moment response of short rectangular Passon, P. 2006. Memorandum derivation and description of the soil-pile
piers in sand.” Comput. Struct. 81 (30–31): 2717–2729. https://doi.org interaction models. IEA-Annex XXIII Subtask 2. Stuttgart, Germany:
/10.1016/S0045-7949(03)00337-7. IEA.

© ASCE 04018141-20 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141


Pender, M. J., D. P. Carter, and S. Pranjoto. 2007. “Diameter effects on pile 17 (1): 04016024. https://doi.org/10.1061/(ASCE)GM.1943-5622
head lateral stiffness and site investigation requirements for pile founda- .0000667.
tion design.” J. Earthquake Eng. 11 (1): 1–12. https://doi.org/10.1080 Sherif, M., Y. Fang, and R. Sherif. 1984. “KA and K0 behind rotating and
/13632460701280005. non-yielding walls.” J. Geotech. Eng. 110 (1): 41–56. https://doi.org/10
Randolph, M. F. 1981. “The response of flexible piles to lateral loading.” .1061/(ASCE)0733-9410(1984)110:1(41).
Geotechnique 31 (2): 247–259. https://doi.org/10.1680/geot.1981.31.2 Sørensen, S. P. H. 2012. “Soil-structure interaction for non-slender large-
.247. diameter offshore monopoles.” Ph.D. thesis, Dept. of Civil Engineering,
Reese, L. C. 1977. “Laterally loaded piles: Program documentation: J. Aalborg Univ.
Geotech. Engng. Div. ASCE Vol. 103 No. GT4 Proc. Paper 12862 Sørensen, S. P. H., L. B. Ibsen, and A. H. Augustesen, 2010. “Effects of di-
(April 1977) pp 287–305.” Comput.-Aided Des. 10 (3): 210. https://doi ameter on initial stiffness of p-y curves for large-diameter piles in sand.”
.org/10.1016/0010-4485(78)90185-9. In Proc., 7th Conf. on Numerical Methods in Geotechnical Engineering,
Reese, L. C., W. R. Cox, and F. D. Koop, 1974. “Analysis of laterally 907–912. Trondheim, Norway: Taylor & Francis.
loaded piles in sand.” In Vol. 2 of Proc., Sixth Annual Offshore Terzaghi, K. 1955. “Evaluation of coefficients of subgrade reaction.”
Technology Conf., 473–485. Dallas: OTC. Geotechnique 5 (4): 297–326. https://doi.org/10.1680/geot.1955.5.4.297.
Reese, L. C., and I. Manoliu, 1973. “Analysis of laterally loaded piles by Thieken, K., M. Achmus, and K. Lemke. 2015. “A new static p-y approach
Downloaded from ascelibrary.org by University Of Kentucky on 08/23/18. Copyright ASCE. For personal use only; all rights reserved.

computer.” Buletinul Stiintific 16 (1): 35–65. for piles with arbitrary dimensions in sand.” Geotechnik 38 (4): 267–
Reese, L. C., and W. R. Sullivan, 1980. Documentation of computer pro- 288. https://doi.org/10.1002/gete.201400036.
gram COM624: Parts I and II, Analysis of stresses and deflections for Vesic, A. B. 1961. “Bending of beams resting on isotropic elastic solid.” J.
laterally loaded piles including generation of p-y curves. Geotechnical Engrg. Mech. Div. 87 (4): 35–51.
Rep. No. GS80-1. Austin, TX: Univ. of Texas at Austin. Wang, S.-T., and W. M. Isenhower. 2010. LPILE, version 6 user’s manual:
Reese, L. C., and W. F. Van Impe. 2011. Single piles and pile groups under A program for the analysis of deep foundations under lateral loads.
lateral loading. 2nd ed. Boca Raton, FL: CRC Press, Taylor and Francis Austin, TX: ENSOFT, Inc.
Group. Wiemann, J., K. Lesny, and W. Richwien, 2004. “Evaluation of pile diameter
Shadlou, M., and S. Bhattacharya. 2016. “Dynamic stiffness of monopiles effects on soil-pile stiffness.” In Proc., 7th German Wind Energy Conf.
supporting offshore wind turbine generators.” Soil Dyn. Earthquake (DEWEK). Wilhelmshaven, Germany: Deutsches Windenergie-Institut.
Eng. 88 (Sept): 15–32. https://doi.org/10.1016/j.soildyn.2016.04.002. Wong, K. S., and J. M. Duncan, 1974. Hyperbolic stress-strain parameters
Sheil, B., and W. Finnegan. 2017. “Numerical simulations of wave- for non-linear finite element analyses of stresses and movements in soil
structure-soil interaction of offshore monopiles.” Int. J. Geomech. masses. Rep. No. TE-74-3. Berkeley, CA: Univ. of California.

© ASCE 04018141-21 Int. J. Geomech.

Int. J. Geomech., 2018, 18(11): 04018141

You might also like