Download as pdf or txt
Download as pdf or txt
You are on page 1of 80

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

net/publication/309174077

Analysis and design of monopile foundations for offshore wind-turbine


structures

Article  in  Marine Georesources and Geotechnology · January 2015

CITATIONS READS

23 6,470

2 authors:

Muhammad Arshad Brendan C. O'Kelly


University of Engineering and Technology, Lahore Trinity College Dublin
23 PUBLICATIONS   271 CITATIONS    191 PUBLICATIONS   2,673 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Water content determinations for peat and other highly organic soils using the oven drying method View project

Geotechnics of biosolids, sewage sludge and water-treatment residue materials View project

All content following this page was uploaded by Muhammad Arshad on 14 December 2018.

The user has requested enhancement of the downloaded file.


Energy Proceedings of the Institution of Civil Engineers
Energy 166 November 2013 Issue EN4
Volume 166 Issue EN4
Pages 139–152 http://dx.doi.org/10.1680/ener.12.00019
Paper 1200019
Offshore wind-turbine structures: Received 22/10/2012 Accepted 08/05/2013
a review
Arshad and O’Kelly Keywords: foundations/offshore engineering/renewable energy

ICE Publishing: All rights reserved

Offshore wind-turbine
structures: a review
g
1 Muhammad Arshad MSc, ME g
2 Brendan C. O’Kelly PhD, FTCD, CEng, CEnv, MICE
PhD candidate, Department of Civil, Structural and Environmental Associate Professor, Department of Civil, Structural and Environmental
Engineering, Trinity College Dublin, Ireland; Lecturer, Department of Engineering, Trinity College Dublin, Ireland
Geological Engineering, University of Engineering and Technology,
Lahore, Pakistan

This paper reviews various issues related to wind-power generation, one of the more popular forms of renewable
energy, including attractions and challenges of electric power generation through onshore and offshore resources.
Significant increases in wind-turbine dimensions, rated power-generation capacity and size of wind farm developments
over the past two decades are projected to continue. Offshore wind-power generation presents many engineering
challenges including: limited guidelines available for analysis and design of foundation/support structures; inadequate
logistics for construction/fabrication; and comparatively expensive operation and maintenance costs, which combined
result in current levelised cost of energy approximately double that for onshore wind-power generation. Different off-
shore foundation options are discussed in terms of general layout, loading characteristics and related fundamental
natural frequency. Outlooks for some new approaches/developments and areas for further research are identified that
may go towards reducing the levelised cost of energy for wind-power generation more in line with that from other
energy resources, thereby enhancing the attractiveness of this industry for potential investors.

Notation generate electricity. Wind turbines are categorised by axis of


A scalar rotation of the main rotor shaft (either horizontal or vertical
EI bending stiffness axis) and whether they are located onshore or offshore (Tong,
fnat first natural frequency 2010). For modern commercial wind turbines, the main rotor
fwind probability density function shaft is horizontally aligned. Rated power-generation capacity
k shape factor quantifying width of wind-speed distribution is mainly dependent on rotor diameter and wind speed
L strut length (IRENA, 2012); for example, if wind speed increases two-fold,
M turbine mass its energy content increases eight-fold. Two key speed terms
U 10 wind speed at 10 m elevation above mean sea level or are ‘cut-in speed’, at which the wind turbine begins to
typically at hub height for OWTs produce power, and ‘cut-out speed’ at which the turbine must
U z mean wind speed at elevation z above mean sea level be shut down to protect the rotor and drive-train machinery
1P first excitation frequency from possible damage (Sørensen et al., 2009; Tong, 2010).
3P blade passing frequency for three-bladed turbine
a scalar Between 2000 and 2011, global wind-power capacity approxi-
m strut mass per unit length mately doubled every 3 years, with an estimated total power
n Weibull random variable generation of 238 GW achieved by the end of 2011; China,
the USA and Germany are the top industry players (GWEC,
1. Introduction 2011). Although the market is still dominated by onshore,
Wind-power turbines harness the kinetic energy of the wind, with significant onshore wind resources yet to be explored, the
providing the motive force to rotate turbine blades and offshore wind market is growing rapidly. Global total installed
develop, by way of a drive shaft, the mechanical power to capacity for offshore of 3.12 GW was generated by the end of

139
Energy Offshore wind-turbine structures:
Volume 166 Issue EN4 a review
Arshad and O’Kelly

2010, with 1.16 GW added in 2010 alone – a 59.4% increase on g trends in geometric size and rated power-generation
the previous year (WWEA, 2011). Total offshore wind-power capacity of onshore and offshore wind turbines
capacity in Europe reached 2.90 GW by the end of 2010, with g cost analyses
0.88 GW added in 2010; again this represents a significant g different foundation options available, including features
increase of 43.6% on the previous year. This occurred at the of exemplary structures, with particular focus on OWT
same time as onshore new-capacity additions declined by 13% structures
(WWEA, 2011). g challenges and attractions of wind-power generation.

The size of offshore wind farms is also increasing, with 2010 Recommendations for future research and practice are also
data indicating that the average size of an offshore wind farm proposed to make offshore wind energy comparable with
in terms of power output was 155 MW – more than double other sources of renewable energy.
the average wind farm size of 72 MW for 2009 (EWEA,
2011). Preliminary data for 2011 suggest that offshore wind- 2. Trends in geometric size and rated power
power capacity in Europe increased by 0.86 GW (EWEA, capacity of offshore wind turbines
2012), with the offshore market likely to be driven by mainly Figure 1(a) shows the main components of an OWT system,
the UK and Germany, although France and Sweden also including a typical monopile foundation, the substructure, tran-
have significant projects imminent. Collectively the European sition piece, tower, rotor blades and nacelle (hub). Modern
Union (EU) has plans to generate approximately 40 GW from OWTs are installed with either pitch-regulated blades or vari-
offshore wind by 2020 (EWEA, 2009). In its 2008 communi- able rotational speed systems in order to allow optimisation
cation on offshore wind energy, the European Commission of power production over a wide range of prevailing wind
anticipated offshore wind can and must make a substantial con- speeds. The rotational speed of the main rotor shaft is typically
tribution to meeting the EU’s energy policy objectives through a between 10 and 20 rpm (Alderlieste, 2010; Malhotra, 2011). The
very significant increase – in the order of 30 to 40 times by 2020 nacelle (Figure 1(b)) contains key electromechanical com-
and 100 times by 2030 – in installed capacity compared to today ponents of the wind turbine, including the gearbox and
(ECN, 2011a). generator. Operational details of these components have been
reported by Maria (2009) and Tong (2010). The gearbox may
Interest in offshore wind power is also increasing in other cause efficiency losses for the wind turbine and is a particular
regions of the world, with, for example, China, the USA and source of noise. Recent developments in the design of perma-
South Korea planning to generate 6.0 and 3.0, 2.5 GW, respect- nent magnet generators have made it possible to construct
ively, by 2020. Building on this, China and the USA have some types of wind turbines without the requirement for a
ambitious plans to generate 65 and 54 GW, respectively, from gearbox. In this case, the rotor is connected directly to a low-
offshore wind by 2030 (AWEA, 2012; Musial and Bonnie, speed multi-pole generator that rotates at the same speed,
2010). termed a direct-drive unit. Removing the gearbox removes
one of the key components requiring more maintenance and
A significant hurdle for the offshore market, however, is the that is prone to failure. This simplification of the mechanical
high initial capital investment costs of the project, which is part allows reductions in size and mass of the nacelle (Treehug-
related to: inadequate and (or) potentially unreliable design ger, 2011).
guidelines for offshore wind-turbine (OWT) installations,
especially foundation structures; more stringent requirements The substructure connects the transition piece or tower to
for durable construction materials to withstand the harsh the foundation at seabed level. In Figure 1(a), a monopile is
marine environment; high-tech equipment requirements for shown as the foundation system, although other foundation
on-site operation and also shortage of trained manpower types, discussed later in the paper, may also be used.
(Musial and Bonnie, 2010). In addition, the next generation of Together the tower, substructure/support structure and foun-
OWTs will be installed at greater distances offshore and hence dation maintain the turbine in its correct operational
in greater water depths (see Section 4). Compared with position. The transition piece provides a means of correcting
onshore, attractions of offshore wind-power generation gener- for any vertical misalignment of the foundation that may
ally include: longer life-span of OWTs on account of less fluctu- have occurred during its installation. In some cases, the foun-
ation of wind speed; availability of ample free space for dation can extend to above the water surface, thereby also
installation; consistently higher wind speeds and generally serving as a substructure by connecting directly to the tran-
reduced adverse environmental effects (Damien and Mo, 2002). sition piece or tower.

This review paper considers the following aspects of the wind Figure 2 shows the steady increase in rotor diameter and rated
industry power capacity (RPC) of wind turbines installed over the past

140
Energy Offshore wind-turbine structures:
Volume 166 Issue EN4 a review
Arshad and O’Kelly

Blade

Nacelle Rotor Tower


diameter Work/service
Hub platform
Intermediate
platform
Sea level
Boat landing Transition
piece Substructure
Tower
Support External J-tube
Work/service structure
platform Scour protection
Sea level
Transition
Seabed level
piece
Substructure
Seabed level
Foundation
Monopile
Monopile
Foundation

(a) (c)

1 6 7 8 9 10 11 12 13 14 15

3
4
5

16 17 18
01. Blade 8. Aircraft warning lights 14. Transformer
2. Blade support 9. Gearbox 15. Anemometers
3. Pitch angle actuator 10. Mechanical brakes 16. Frame of the nacelle
4. Hub 11. Hydraulic cooling devices 17. Yaw driving device
5. Spinner 12. Generator 18. Supporting tower
6. Main support 13. Power converter and electrical control,
7. Main shaft protection and disconnection devices
(b)
Figure 1. Major components of OWT system: (a) wind-turbine
system; (b) electromechanical parts adapted from
ABB (2012); (c) details of monopile and transition piece

three decades. In particular, between 1990 and 2010, the RPC phase (EWEA, 2011b). Table 1 presents correlations deter-
increased from typically 0.5 to 7.5 MW and rotor diameter mined from data of more than 150 modern, utility-scale wind
from approximately 40 to 150 m (EWEA, 2011b). Offshore turbines which can be used to approximate the size and mass
wind turbines having 250 m rotor diameters and with of different OWT components, considering RPC as a key
RPC ≥ 20 MW are currently in the research and development driving input.

141
Energy Offshore wind-turbine structures:
Volume 166 Issue EN4 a review
Arshad and O’Kelly

300

250

20 MW
Rotor diameter: m

200

6 MW
150 10 MW
7·5 MW
100
0·1 MW 5 MW
0·5 MW
50
0·01MW
0·6 MW
0
1975 1980 1985 1990 1995 2000 2005 2010 2015 2020 2025
Year
Figure 2. Increase in rotor diameter and RPC of wind turbines

3. Cost analysis of wind-power generation Between 40% and 70% of costs for conventional fossil-fuel-fired
Approximately 70–75% of the total cost of offshore wind-power technologies are related to fuel, operation and maintenance
production is related to initial capital investment costs, including (Søren et al., 2009). Hence, since fuel costs have no impact on
that of the turbine, foundation, electrical equipment and grid wind-power generation costs, wind turbines are more capital
connection (Kooijman et al., 2001; Søren et al., 2009). The intensive compared with fossil-fuel-fired technologies. In China
‘levelised cost of energy’ (LCOE) is the primary measure for for instance, the LCOE for onshore wind was almost 300%
quantifying and comparing underlying economics of power and 200% more costly compared with electric power generation
projects (Fischer, 2011). For wind-power systems, LCOE from natural gas and coal, respectively (YFH, 2011), although
represents the sum of all costs, including capital cost, operation such cost comparisons are somewhat dependent on the accuracy
and maintenance costs, and also expected annual energy of projected trends for the costs of fuel, other commodities and
production (Cambell, 2008; Søren et al., 2009) for a fully logistic facilities. Initial capital investment costs for offshore are
operational wind-power system over the project’s lifetime, with approximately double (YFH, 2011) and may reach up to three
financial flows discounted to a common year. However, empirical times that for onshore wind-power projects having similar
methods that use the more extensive databases currently avail- power generation capacity on account of increased investment
able for onshore wind-power projects in estimating the LCOE required in transportation of materials and turbines, construction
for new offshore projects are not reliable (IRENA, 2012). and installation of foundations, equipment and turbines at sea

Parameters Correlation

Rotor diameter (D, in m) and RPC (MW) D = 59.354(RPC)0·47


Rotor speed (Rspeed, in rpm) and RPC (MW) Rspeed = 22.781(RPC)−0·3595
Hub height (HH) and rotor diameter (D) HH = D/0.255(D)0·3464
Hub mass including pitch, bearing and driver system (M(pb + ds), in t) and RPC (MW) M(pb + ds) = 8.6421(RPC)1·1194
Rotor mass including hub, pitch system and blades (M(hb + ps + bl), in t) and RPC (MW) M(hb + ps + bl) = 18.453(RPC)1·1357
Mass of main rotor shaft (M(ms), in t) and RPC (MW) (M(ms)) = 0.2415(RPC)2 + 3.0699(RPC)
Mass of main bearing (M(mb), in t) and RPC (MW) M(mb) = 0.1246(RPC)2 + 1.2623(RPC)
Mass of rotor, drive-train support structure and nacelle (M(r + d + n), in t) and RPC (MW) M(r + d + n) = 37.45(RPC)0·984
Mass of all components at top of tower (M(thm), in t) and RPC (MW) M(thm) = 55.9216(RPC)1·0341
Table 1. Effect of RPC on size and mass of wind-turbine
components (Tong, 2010)

142
Energy Offshore wind-turbine structures:
Volume 166 Issue EN4 a review
Arshad and O’Kelly

and laying offshore cables (IRENA, 2012; Martin et al., 2004). also be necessary to ensure smooth operation of the next gener-
The trend for OWT installations at increasing distances offshore ation of OWTs to be installed at greater distances from the
and hence location in greater water depths constitutes a signifi- shore line. Hence it is clear that the reduction of O&M costs for
cant factor in the cost analysis for offshore projects. Cost offshore wind farms is a key challenge, and once addressed,
comparisons between onshore and offshore wind-power technol- may improve the economics of offshore wind energy (Douglas-
ogies should be based on evaluations for a specific region and/or Westwood Limited, 2002; IRENA, 2012).
country (EEA, 2009). Table 2 shows comparisons between costs
for different components of onshore and offshore wind-energy In Europe, LCOE estimates of between 0.10 and 0.13 US$/kWh
projects. were reported by IEA (2009) for onshore wind in 2011, assum-
ing a typical capacity factor (ratio of average power delivered to
For wind-power generation, the overall contribution of operation theoretical maximum power) for new onshore projects of
and maintenance (O&M) costs to the LCOE is significant and also between 25% and 30%. For a given capacity factor, assumed
site specific. Data from different countries including the USA, cost reductions achievable by 2015 may allow reductions in
China and many European countries indicate that O&M costs LCOE of between 6% and 7%. In North America, the LCOE
for onshore wind power account for between 11% and 30% of for onshore wind having a capacity factor of 30% was estimated
the total LCOE (IRENA, 2012). The lowest contribution of at between 0.10 and 0.11 US$/kWh for 2011. By 2015,
US$0.010/kW was reported for the USA, with approximately anticipated cost reductions for a given capacity factor may
US$0.013-0.015/kW reported for best practice in Europe allow reductions in LCOE of between 5% and 9% (Wiser and
(IRENA, 2012). However, O&M costs for offshore are signifi- Bolinger, 2011). Compared with Europe and North America,
cantly greater on account of higher costs incurred in accessing LCOE estimates for onshore wind power in China and India
and maintaining the wind turbines, towers and cabling. In the were significantly lower at between 0.07 and 0.08 US$/kWh
UK, for example, Feng et al. (2010) reported O&M costs for (2011 data) for a capacity factor of 25%. However, since
offshore wind-power projects in shallow water depth were China and India already have very competitive installation
approximately 1.5 times that for onshore projects. Offshore main- costs for wind-power projects compared to the norm in other
tenance costs are also higher on account of the harsher marine developed countries, opportunities for further cost reductions
environment and higher expected failure rates for some electrical are comparatively smaller. By 2015, average installation costs
and mechanical components. In general, O&M costs for offshore may also increase somewhat on account of projected increases
wind power are typically in the range US$0.027–0.054/kWh in engineering project costs, manufacturing costs for wind
(ECN, 2011b). Many existing offshore wind farms are only at turbines in emerging economies and/or the supply situation
the beginning or early stage of their deployment phase. Since becoming tighter (E.ON Climate & Renewables, 2011).
data on their O&M costs remain highly project specific, it will
be some time before observable trends emerge and means of redu- As a general trend, the LCOE for offshore wind-power gener-
cing these costs are identified. Offshore maintenance facilities may ation around the globe is typically almost double that of

Item Offshore Onshore

Initial capital investment cost: US$/kW 3300–5000 1700–2450


Wind-turbine cost, including production, transportation and installation: % of initial capital investment 30–50 65–84
cost
Cost of grid connection including cabling, substations and buildings: % of initial capital investment 15–30 9–14
cost
Construction cost including foundation, transportation and installation of tower and turbine and other 15–25 4–16
infrastructure (e.g. access roads for onshore) necessary for turbine installation: % of initial capital
investment cost
Other capital costs including development and engineering costs, licensing procedures, consultancy, 8–30 4–10
permits, supervision, control and data acquisition, monitoring systems: % of initial capital investment
cost
Table 2. Cost comparisons for offshore and onshore wind projects
(Douglas-Westwood Limited, 2002; Henderson et al., 2003;
IRENA, 2012; Junginger et al., 2004; Kooijman et al., 2001)

143
Energy Offshore wind-turbine structures:
Volume 166 Issue EN4 a review
Arshad and O’Kelly

onshore having similar capacity factors (Roddy et al., 2009).


For instance, reported ratios of LCOE for offshore to
onshore wind-power projects were 1.3 for Denmark and 1.46 Ambient turbulence
for the UK (Feng et al., 2010; Krohn et al., 2009). Between Wake turbulence
Operational and
2009 and 2011, the overall trend in the LCOE for offshore accidental loads
wind power continued to increase gradually, compared with
onshore, which typically showed a small reduction (BNEF,
2011). Hence the LCOE for offshore wind power is likely to
remain greater than for onshore (even taking into account Icing
Marine
higher capacity factors achievable offshore) and will probably growth
remain so, given the significant challenges involved in reducing
capital and O&M costs (Tricklebank, 2008). The main reason Waves
Ship impact
for this is the trend of increasing distance offshore (and hence
greater water depth) necessary for the next generation of off-
Currents
shore wind-farm operations, which leads to increased costs in
all aspects of the supply chain. Turbine prices are increasing Seabed
Scour
due to necessary design improvements to achieve greater
reliability in the harsh marine environment and also on
account of larger, more sophisticated wind turbines necessary
Figure 3. Environmental impacts on OWTs
to increase capacity factors (Maria, 2009). Construction and
cabling costs also increase as a function of water depth/distance
offshore. However, encouragingly, a recent study performed in
China (IEA, 2011) indicated that by 2020, 2030 and 2050, initial offshore wind farms located in shallow water depths (≤40 m).
capital investment costs for offshore wind-power projects Such monopiles consist of a steel tubular section (pipe pile),
located in up to 50 m water depth are estimated to reduce by typically 4–6 m in diameter and up to 1000 t in mass, which trans-
25%, 36% and 50%, respectively, compared with current fers the applied vertical and larger lateral loading into the seabed
initial capital investment costs (2010 data). Furthermore a foundation. Its complete installation is usually achieved within
33% reduction in O&M costs was predicted by 2030, although 24 h (Fischer, 2011; Junginger et al., 2004; Saleem, 2011).
no further reduction was anticipated between 2030 and 2050. In Cyclic lateral and moment loading on the monopile are resisted
Europe, respective reductions of 15% and 20% in initial capital by horizontal earth pressures mobilised in the surrounding soil
investment and O&M costs are estimated to occur between 2006 along the monopile embedded length. The monopile embedment
and 2015, although further reductions of approximately 10% in length is dependent on seabed characteristics/properties and total
initial capital investment costs are expected for offshore wind- applied load. An embedment length of 30 m is usually deemed
power projects between 2020 and 2050 (CEC, 2007; Søren sufficient to meet design criteria, including vertical stability and
et al., 2009). horizontal deflection requirements (Musial and Bonnie, 2010;
Tricklebank, 2008).
4. Foundation systems
Support structures/foundations for offshore wind farms are Braced support structures (i.e. tripod and jacket/truss) are more
generally more complex than for onshore, involving greater suitable for deeper water and heavier turbines (Esteban et al.,
technical challenges, including design requirements to with- 2011; Fischer, 2011). Tripods consist of a large-diameter
stand the harsh marine environment and prolonged impact central steel tubular section that is supported over its lower
under large wave loading (see Figure 3). The various support length by three braces (Figure 4(d)), which are connected to
structure/foundation concepts employed (Figure 4), which the seabed. A range of different foundations can be employed
have been adopted from the offshore oil and gas industry, are including gravity bases, suction buckets or piles. In this
usually categorised as either bottom-mounted structures (i.e. manner, the structural and environmental loads applied on
rigidly connected to the seabed through a foundation system) the OWT and the supporting structure are mostly transferred
or floating-support structures that have no rigid connection axially through the braces to the seabed foundation. Complete
with the seabed. The foundation solution adopted depends on installation of a tripod foundation system with, for example, a
local seabed conditions, water depth and financial constraints water surface to seabed length of up to 50 m and mass of up
(AWS True wind, 2010; Igoe et al., 2013). to 700 t, typically takes between two and three working days,
often requiring special equipment for driving/drilling and
At present, monopiles (Figures 4(b) and 4(c)) are by far the most working under water (Esteban et al., 2011; Fischer, 2011). A
widely adopted substructure–foundation system for modern jacket/braced frame structure (Figure 4(e)) is a lattice frame

144
Energy Offshore wind-turbine structures:
Volume 166 Issue EN4 a review
Arshad and O’Kelly

Sea level
0

10

20 (a)
Water depth: m

30
Seabed
40

50
(b)
60 (c)

70
(d)
80
(e)

(f) (g)

Figure 4. Support structure/foundation options for OWTs:


(a) gravity; (b) monopile; (c) monopile with guy wire; (d) tripod;
(e) braced frame; (f ) tension leg with suction buckets (ballast stabilised);
(g) buoy with suction anchor

of small-diameter steel struts that (similar to tripods) is and rotation around the vertical axis (yaw). Other examples
anchored to the seabed using one of the different foundation include spar-floater and barrage-floater systems. For the spar-
options. Complete installation generally takes up to 3 days. floater (Figure 4(g)), buoyancy is provided to the wind-turbine
Braced frame structures are particularly suitable for severe structure by a long, slender cylinder/capsule that protrudes
maritime weather since the strut components offer lower resist- below the water line (Esteban et al., 2011; Fischer, 2011; Vries,
ance to prevailing ocean wave and current flow in comparison 2007). For the barrage floater, the wind-turbine structure is
with monopile or tripod structures. Braced frame structures placed on a barrage and attached by way of anchor lines to the
are also more adaptable to conditions encountered on site, seabed.
increasing their application range, with geometrical variations
of the substructure achieved relatively simply but without alter- From the various foundation systems described above, mono-
ing the stiffness of the whole structure (Vries, 2007). piles are currently by far the most popular solution used
worldwide, with 75% share, in comparison with only 5% for
In the future, it is anticipated that floating structures, which are jacket/tripod options (E.ON Climate & Renewables, 2011).
currently only at research and development stage, will be com- However, it is estimated that by 2020, between 50% and 60%
mercially used, particularly for water depths greater than 50 m of new OWTs will be supported by monopiles and a further
(Saleem, 2011). Such floating platforms for wind turbines will 35–40% by jacket/tripod systems (Babcock and Brown
impose many new design challenges. Currently, tension-leg Company, 2012). The main reason for this shift is the attraction
platform concepts (see Figure 4(f)) are considered as most of jacket/tripod systems for deeper sea locations, which provide
economical (Fischer, 2011) because rigid body modes of the consistently higher wind speeds and hence greater wind energy
floater are limited to horizontal translation (surge and sway) (Tempel and Molenaar, 2002).

145
Energy Offshore wind-turbine structures:
Volume 166 Issue EN4 a review
Arshad and O’Kelly

5. Comparison of environmental loading for probability density function ( fwind) of a Weibull random vari-
offshore and onshore wind-turbine able n is given by
structures      
k n k−1 n k
Offshore wind-turbine structures are designed to resist loading fwind = exp −
from hydrodynamic, aerodynamic and also ice and ship-impact 2. A A A
sources, whereas onshore structures are principally designed to
withstand aerodynamic loading. Aerodynamic loading results where A is a scalar and k is a shape factor that quantifies the
from interactions of the rotor and parts of the tower within the width of the wind-speed distribution.
turbulent air field, with the generated wind power directly pro-
portional to the cube value of mean wind speed. Aerodynamic The values of A and k are larger for offshore, indicating higher
conditions for offshore and onshore scenarios are markedly probability of greater wind speeds compared with onshore
different, with considerably lower fluctuation in loading (Fischer, 2011; Tricklebank, 2008). Greater differences in
experienced for offshore on account of associated free-flow con- wind-speed distributions over time for offshore compared
ditions and lower surface roughness, although advantages of with onshore produce higher levels of mean wind load and
reduced dynamic loading are partly undone by higher mean hence greater power output. Long-term variations in wind
wind speeds (Fischer, 2011; Tricklebank, 2008). In general, speed are significant in terms of predicting energy yield from a
aerodynamic loading can be characterised by (DNV, 2011) wind turbine, whereas short-term fluctuations are more relevant
for generated wind loads. The degree of turbulence (defined as
g vertical wind profile momentary deviations from the mean wind speed) depends on
g mean wind-speed distribution meteorological and geographical conditions, for example,
g turbulence effects. atmospheric layering and terrain. The main contributors to
extreme loading and fatigue are stochastic effects in short-
For offshore, surface roughness is low, increasing only margin- term fluctuations of wind speed, such as turbulence/transient
ally in the event of severe sea states with high waves. Hence wind events such as gusts (Quarton et al., 1996). A measure for turbu-
speed increases sharply with increasing elevation above sea lence is turbulence intensity: the ratio of the standard deviation
level, producing very steep wind-speed profiles compared with of wind speed to mean wind speed for a given time period
onshore sites. The mean value of 10 min wind-speed data (DNV, 2011). For a particular site, turbulence intensity corre-
(either measured at a reference elevation of 10 m above mean lates with wind speed and surface roughness; higher wind
sea level or usually determined at hub height for OWTs) is speed and lower surface roughness produce lower turbulence
referred to as wind speed U 10 . The mean wind speed U z at (Vries, 2007). Since wind is the primary energy source for
some other elevation z above mean sea level can be approxi- ocean waves, higher wind speed may produce marginal
mated by increases in turbulence on account of ensuing increases in
 a roughness of the ocean surface (Letchford and Zachry, 2009).
z Another aspect of fluctuating wind speed is turbulence
U z = U 10
1. 10 induced in wake conditions (see Figure 3). Ambient non-
obstructed turbulence is the ‘normal’ turbulence experienced
where values of a range between 0.11 and 0.40 depending on site by a single stand-alone turbine at a particular site (Frandsen
location; for example, a = 0.11 for open sea conditions, 0.16 for and Thøgersen, 1999). Wake effects can be significant, especially
grassland and 0.40 for city centre/urban environments (Haritos, for dense wind-park layouts, where neighbouring turbines
2007; Journée and Massie, 2001). experience a superimposed turbulent wind coming from the
ambient and wake.
For offshore sites, steep profiles of wind speed for the vertical
direction usually necessitate lower hub heights, with minima Since surface roughness and hence ambient turbulence are lower
values generally dictated by clearance limits to the turbine’s for offshore sites, the combination of ambient and wake-
service platform (see Figure 1(a)). Periodic loading effects are induced turbulence is also comparatively lower although wake
also reduced since the difference in mean wind speed between fields remain longer in the atmosphere compared with
upward and downward moving blades is low (Fischer, 2011). onshore. The frequency of energy-rich wind turbulence is
In contrast, the gain in wind energy with increasing hub below 0.1 Hz (LeBlance, 2009). Hence, turbulence is not
height is the driver for onshore design. Wind-speed distribution, significant in the determination of the structural design loads
which also differs between onshore and offshore, is generally for extreme levels of environmental loading, although its
described by a Weibull distribution function (DNV, 2011) effect on fatigue life of wind-turbine structures cannot be
that quantifies the probability of different mean wind speeds ignored (Vries, 2007). Offshore wind turbines are generally
occurring over a given time period at the site location. The designed for more severe wind classes since the probability of

146
Energy Offshore wind-turbine structures:
Volume 166 Issue EN4 a review
Arshad and O’Kelly

extreme wind speeds (e.g. due to gusts or changes in wind this invariably leads to the development of higher stresses in
direction) is more significant compared with most onshore the support structure and foundation, but more significantly
sites (GL, 2005). to a higher range of stresses – an unfavourable situation in
considering fatigue life. Hence it is important to ensure that
Compared with the rotor thrust reaction to wind loads, the excitation frequencies having high energy levels do not coincide
hydrodynamic forces acting on OWTs generally only have a with the support structure’s fundamental natural frequency. As
minor role in the development of tower deflection. LeBlance a first approximation, the support structure’s fundamental
(2009) reported that this was largely due to the reduced wave-in- natural frequency can be determined by considering a simplified
teraction area of the substructure as compared with the overall geometry for the whole structure (Figure 5). The turbine mass
tower length and greater lever arm of the rotor thrust (see (M) is concentrated at the top (free end) of an equivalent steel
Figure 1(a)). However, the density of the medium must also pipe representing the support structure, with similarities to a
be considered when comparing aerodynamic and hydrodynamic cantilevered vertical strut. In this instance, the first natural
lateral loading, with the density of sea water significantly greater frequency ( fnat , in Hz) of the combined structure can be
than that for air. Hydrodynamic forces generally only become approximated by (Tempel, 2006)
significant for greater water depths and/or wave heights,
which cause the lever arm of the hydrodynamic force to increase 3.04 EI
2
fnat =  
along with the intensity of the lateral force generated by the 3. 4(p)2 0.227mL + M L3
water (Fischer, 2011). The height of the ocean waves is
usually expressed in terms of ‘significant wave height’; this is where m is the strut mass per unit length, L is the strut length
defined as the mean value of the highest one-third of the and EI is its bending stiffness (N m2).
waves in a given wave record. Ocean waves that induce
fatigue loading with high frequency usually have significant Offshore wind-turbine structures are excited by both wind and
wave heights of 1.0–1.5 m and a zero-crossing period of waves, with the effective wind load determined by complex
4–5 s (Vries, 2011). interactions between the structural dynamics of the turbine
and wind field. Site-specific spectral densities for wind and
6. Loading frequency, natural vibration waves can be derived either from data measured for the particu-
frequency and resonance lar site location, from met-ocean databases or numerical models
It is essential to consider the fundamental natural frequency of a (LeBlance, 2009).
wind-turbine structure for a proper description and evaluation
of its dynamic behaviour. As for all dynamic systems, resonance Dynamic amplification and large excitation forces affect mono-
occurs when an excitation frequency gets close to the structure’s piles in a cumulative unfavourable manner. With rotational
fundamental natural frequency. For wind-turbine structures, speeds of the main rotor shaft typically between 10 and
20 rpm, the first excitation frequency ‘1P’ (i.e. corresponding
to one full revolution) occurs in the range 0.17–0.33 Hz. In
general, only light excitation of the 1P frequency should
M, turbine mass occur, with large excitations arising on account of excessive
mass and/or aerodynamic imbalances. For a three-bladed
turbine, the blade passing frequency of typically 0.5–1.0 Hz is
µ, strut mass per unit length

EI, strut bending stiffness Wind turbulance


L, length

Waves
Power spectral density

Wanted
1P frequency 3P

0·0 0·1 0·2 0·3 0·4 0·5 0·6 0·7 0·8 0·9 1·0 1·1
Frequency: Hz
Figure 5. Structural model for flexible wind-turbine system Figure 6. Excitation ranges for OWT structures

147
Energy Offshore wind-turbine structures:
Volume 166 Issue EN4 a review
Arshad and O’Kelly

denoted as the ‘3P’ frequency and is heavily excited on account 7. Challenges for offshore wind-power
of impulse-like excitation arising from the individual blades generation
passing by the tower. Offshore wind-power generation has arguably greater potential
compared with onshore, but marine conditions pose great
Figure 6 shows the excitation ranges of 1P and 3P, along with challenges to project delivery because outcomes are highly
realistic, normalised power spectra for wind and wave exci- influenced by environmental conditions. Some outlooks and
tations. Referring to the figure, the ‘soft–stiff ’ zone includes new approaches to offshore windpower are discussed below.
the 1P, 3P and ‘wanted frequency’ regions; the region before
the 1P range is referred to as the ‘soft–soft’ zone and the 7.1 Implementation, fabrication, operation and
region after the 3P range as ‘stiff–stiff ’. ‘Soft–soft’ and ‘stiff– maintenance
stiff ’ zones are unsuitable for the design solution. The structure It could be argued that, owing to the involvement of multiple
is considered too flexible if its fundamental natural frequency regulatory and planning bodies, the current planning and
falls within the ‘soft–soft’ zone and too rigid (heavy and expens- implementation process for offshore wind-power farms is too
ive) within the ‘stiff–stiff ’ zone. Another important reason for complex and time-consuming. Some primary legislation may
avoiding the ‘soft–soft’ frequency region is that wave- and be helpful in order to facilitate relevant government bodies
wind-turbulence excitation frequencies usually fall within this working amicably with investors and developers for offshore
zone (LeBlance, 2009); see Figure 6. wind-power farms (CT, 2008).

Excitation/resonance of a dynamic system can be mitigated by Fabrication and O&M issues place major pressures on the LCOE
damping, achieved either internally by friction in components for offshore wind farms (ABB, 2012; IRENA, 2012). New off-
of the structural system or externally by some source/force. shore strategies must be developed to minimise the number of
Overall damping of an offshore structure can be achieved by tasks performed at offshore sites. Materials for wind-turbine
combinations of aerodynamic, hydrodynamic, structural and/ fabrication must be selected for durability and environmental
or soil damping (refer to Bittkau (2010), Genta (1998) and tolerance. Engineering design, beginning from preliminary con-
Rodenhausen (2010) for further details). cepts, must rigorously place higher premiums on reliability,
float-out deployments and in situ repair methods. Fabrication
For the popular monopile foundation systems, ‘soft–stiff ’ design facilities must be strategically located for mass production,
necessitates relatively high structural and dynamic stiffness, onshore assembly and rapid deployment offshore, with
which can be achieved by increasing the monopile diameter, minimal dependence on large vessels (EWEA, 2007). Sensitive
or less efficiently, by increasing (reinforcing) the pile wall electronic devices for remotely sensing weather conditions and
thickness. However, larger diameter monopiles introduce draw- self-diagnostic systems to manage O&M of electromechanical
backs, including greater wave loading and also larger driving components are required in order to minimise downtime and
equipment/forces necessary for installation (Schaumann and reduce equipment necessary for repairs (CleanTech, 2012). Ulti-
Böker, 2005). Hence there is a corresponding increase in the mately a new balance between initial capital investment costs
initial capital investment costs of the project, although from and long-term operating costs needs to be established that will
the authors’ perspective the LCOE may not be adversely have a broad impact on the LCOE for offshore wind technology.
affected if rated power capacity is also increased by using
larger rotor blades. Compared with monopiles, the lattice 7.2 Offshore design codes and methods
frame of jacket/truss support structures (Figure 4(e)) provides One of the immediate challenges for design is the ability to
large structural bending stiffness and more favourable accurately predict the magnitude and distribution of applied
mass-to-stiffness ratio, resulting in relatively high bending environmental loads and the resulting dynamic response of the
Eigen-frequencies and reduced hydrodynamic excitation coupled wind-turbine and support structure under the action
(Vries, 2011), although torsional stiffness is reduced, potentially of combined stochastic wave and wind loading (Musial and
leading to dynamic problems. Jacket support structures are Butterfield, 2006). At present, analysis, design and installation
designed for operation in or around ‘stiff–stiff ’ regions of monopile foundations for wind-turbine structures usually
(Fischer, 2011). In the case of tripods, bracing along the lower rely on general geotechnical standards, complemented by
length of the central tubular section increases overall bending more specific guidelines and semi-empirical formulas developed
stiffness and reduces bending moment loading on the foun- by the offshore oil and gas industry (API, 2007; DIN, 2005;
dation (Saleem, 2011; Schaumann and Böker, 2005), with DNV, 2011; GL, 2005).
typical Eigen-frequencies ranging between those for monopile
(at lower end of this range) and jacket support structures/ However, large-diameter monopile foundations for proposed
foundations under similar rotor–nacelle configurations and offshore wind-turbine structures are well outside the scope of
environmental conditions. current experience and analysis/design methods (including the

148
Energy Offshore wind-turbine structures:
Volume 166 Issue EN4 a review
Arshad and O’Kelly

American Petroleum Institute (API, 2007) and Det Norske- 8. Summary and conclusions
Veritas (DNV, 2011) standards used by the offshore oil and Wind power, particularly from offshore turbines mounted on
gas industry). These standards are largely based on empirical bottom support structures, appears to be a promising solution
data obtained for relatively small-diameter flexible piles under to meet the universal demand for clean, cost-effective energy.
low numbers of load cycles. Furthermore, for these standards, The rated power-generation capacity of individual OWTs, and
wave loading is of primary concern when extrapolating to also of wind farms, has increased many-fold over the past two
predict extreme events. However, designers of offshore wind- decades, with strong growth projected to continue for the near-
turbine structures must consider wave and wind load spectra to-medium future. Concepts for floating foundation systems for
simultaneously (IEC, 2005; Tarp-Johansen, 2005). Hence OWT structures are also emerging, which will allow installations
careful consideration of these differences in applied loading even farther offshore, thereby benefiting from relatively higher
and inherent limitations underpinning semi-empirical formulae wind speeds/power generation. However, compared with conven-
for the offshore oil/gas industry are required in extrapolation tional fossil-fuel-fired technologies, initial capital investment
for design of large-diameter monopile foundation supports for costs and LCOE for wind-power generation are both compara-
wind-turbine structures. Often these formulations cannot be tively higher, particularly for offshore, on account of challenges
applied with confidence by the offshore wind-power industry associated with the harsh marine environment. Reductions in
to achieve optimum results and economy (Dobry et al., 1982). the LCOE and projected increases in design life of OWT projects
are achievable by developing and implementing improved design
There is also a dearth of knowledge concerning the behaviour of criteria/methods for foundations, support structures and the
the monopile–soil foundation system and its structural stability wind turbines themselves, along with the use of innovative
under long-term cyclic lateral loading. Existing literature materials in their fabrication.
includes Matlock (1970), Reese et al. (1974, 1975), Little and
Briaud (1988), Ismael (1990) and Long and Vanneste (1994). The wind-power generation industry can be facilitated through
Hence the development of more realistic strain-accumulation legislation leading to primary reforms in the rules/regulations
models and also computer codes to predict dynamic forces imposed by different regulatory and monitoring bodies related
and resulting displacements of OWTs will provide valuable to this industry. A multidisciplinary and integrated approach is
tools for more reliable designs. required, with cost reductions achievable for other offshore
industries (e.g. oil/gas sector and offshore cable laying) poten-
7.3 Recommendations for future research and practice tially also benefiting offshore wind-power projects, although
A multidisciplinary approach is suggested in order to strive developments in commodity prices (particularly steel, copper,
towards making the offshore wind-power industry more cement) will also influence potential cost reductions achievable
economical and practicable, including the following points. for wind power. In spite of such challenges, it is projected that
wind-power generation will continue to increase many-fold,
g Some primary legislation may be helpful in order to particularly in Europe, North America and Asia over the next
facilitate relevant government bodies in working amicably two decades, with associated LCOE anticipated to become
with investors and developers for offshore wind-power comparable with other sources of renewable energy.
farms. Investors and developers for offshore-wind projects
could be facilitated by government with some easing of Acknowledgements
current requirements to obtain obligation certificates. The first author gratefully acknowledges a postgraduate
g Development of innovative fabrication materials (with research award from Trinity College Dublin.
appropriate strength, durability and lightweight
characteristics) for OWTs may contribute to considerable REFERENCES
reductions in the LCOE. ABB (2012) Wind Power Plants. Technical Application Paper
g Use of more sophisticated electromechanical parts in No. 13. See http://www05.abb.com/global/scot/scot209.nsf/
OWTs (e.g. direct-drive units that eliminate the veritydisplay/92faf0c1913f5651c1257937002f88e8/$file/
requirement for a gearbox, thereby removing one of the 1sdc007112g0201.pdf (accessed 10/10/2012).
key components prone to failure) will increase the Alderlieste EA (2010) Experimental Modelling of Lateral
efficiency and hence energy yield and also reduce O&M Loads on Large Diameter Mono-pile Foundations in Sand.
costs for the project. MSc thesis, Delft University of Technology, the
g In-depth experimental and numerical studies are necessary Netherlands.
in order to bridge the knowledge gap between existing API (American Petroleum Institute) (2007) API RPA2:
design codes/guidelines developed for the offshore oil and Recommended practice for planning, designing and
gas industry and more onerous applied loading and larger constructing fixed offshore platforms – working stress
support structures/foundations adopted for OWTs. design, 22nd edn. API, Washington, DC, USA.

149
Energy Offshore wind-turbine structures:
Volume 166 Issue EN4 a review
Arshad and O’Kelly

AWEA (American Wind Energy Association) (2012) Facts webarchive.nationalarchives.gov.uk/+/http://www.dti.gov.


About Offshore Wind. See http://offshorewindworks.tumblr. uk/energy/renewables/publications/pdfs/offshorereport.pdf
com/facts (accessed 18/09/2012). (accessed 05/10/2012).
AWS True wind (2010) New York’s Offshore Wind Energy ECN (Energy Research Centre of the Netherlands) (2011a)
Development Potential in the Great Lakes: Feasibility Study; Roadmap to the Deployment of Offshore Wind Energy in the
2010. See http://www.awstruepower.com/wp content/ Central and Southern North Sea (2020–2030). See http://
media/2010/09/NYSERDA_AWST_NYGreatLakesFS.pdf www.windspeed.eu/media/publications/WINDSPEED_
(accessed 10/07/2012). Roadmap_110719_final.pdf (accessed 02/03/2013).
Babcock and Brown Company (2012) The Future of Offshore ECN (2011b) Properties of the O&M Cost Estimator. See http://
Wind Energy. See http://bluewaterwind.com (accessed www.ecn.nl/docs/library/report/2011/e11045.pdf (accessed
20/08/2012). 04/02/2013).
Bittkau R (2010) Concept Study for Using Active Turbine EEA (Eurpean Environmental Ageny) (2009) Europe’s Onshore
Controls vs. Structural Control Systems for Load Mitigation for and Offshore Wind Energy Potential: An Assessment of
Offshore Wind Turbines Using Monopile Support Structures. Environmental and Economic Constraints. See http://www.
University of Stuttgart, Stuttgart, Germany, Research Report. envirocentre.ie/includes/documents/Europes_onshore_and_
BNEF (Bloomberg New Energy Finance) (2011) Levelised Cost offshore_wind_energy_potential[1].pdf (accessed 20/01/2013).
of Energy Update, Q2 2011, Research Note. BNEF, E.ON Climate & Renewables (2011) Offshore Wind Energy
London, UK. Fact Book. See http://www.eon-einkauf.com/content/dam/
Cambell M (2008) The Drivers of the Levelised Cost of eoncom/en/downloads/e/EON_Offshore_Wind_Factbook_
Electricity for Utility-scale Photovoltaics. See http://large. en_December_2011.pdf (accessed 16/08/2012).
stanford.edu/courses/2010/ph240/vasudev1/docs/sunpower. Esteban M, Lopes-Gutierrez J, Diez J and Negro V (2011)
pdf (accessed 15/08/2012). Foundations for offshore wind farms. Proceedings of the
CEC (Commission of the European Communities) (2007) 12th International Conference on Environmental Science and
Renewable Energy Road Map, Renewable Energies in the Technology, Rhodes, Greece, pp. 516–523.
21st Century: Building a More Sustainable Future. See EWEA (European Wind Energy Association) (2007) Delivering
http://ec.europa.eu/energy/energy_policy/docrenewable_ Offshore Wind Power in Europe. See http://www.ewea.org/
energy_roadmap_full_impact_assessment_en.pdf (accessed fileadmin/ewea_documents/images/publications/
10/09/2012). offshorereport/ewea-offshore_report.pdf (accessed 10/10/
CleanTech (2012) Offshore Wind: the Technology Challenge. 2012).
See http://www.cleantechinvestor.com/portal/wind-energy/ EWEA (2009) Pure Power Wind Energy Targets for 2020 and
5573-offshore-wind-the-technology-challenge.html (accessed 2030. See www.ewea.org/fileadmin/ewea.../Pure_Power_III.
10/08/2012). pdf (accessed 14/03/2012).
CT (Carbon Trust) (2008) Offshore Wind Power: Big EWEA (2011a) The European Offshore Wind Industry Key
Challenge, Big Opportunity Maximising the Environmental, Trends and Statistics 2010. See http://www.ewea.org/
Economic and Security Benefits. See http://www. fileadmin/files/library/publications/statistics/EWEA_stats_
carbontrust.com/media/42162/ctc743-offshore-wind-power. offshore_2011_02.pdf (accessed 06/08/2012).
pdf (accessed 10/07/2012). EWEA (2011b) Design Limits and Solutions for Very Large
Damien T and Mo M (2002) The Potential for Renewable Wind Turbines. UpWind Project. See http://www.ewea.org/
Energy Usage in Aquaculture. See http://www. fileadmin/ewea_documents/documents/upwind/21895_
aquacultureinitiative.eu/Renewable%20Energy%20Report. UpWind_Report_low_web.pdf (accessed 04/02/2013).
pdf (accessed 05/08/2012). EWEA (2012) The European Offshore Wind Industry Key
DIN (Deutsches Institut für Normung) (2005) Baugrund Trends and Statistics 2011. See http://www.ewea.org/
Sicherheitsnachweise im Erd und Gundbau. DIN, Berlin, fileadmin/files/library/publications/statistics/EWEA_stats_
Gemerany, DIN 1054: 2005. See http://www. offshore_2011_02.pdf (accessed 06/08/2012).
baunormenlexikon.de/Normen/DIN/DIN%201054/1b69b621- Feng Y, Tavner PJ and Long H (2010) Early experiences with
74a0-4c33-8b29-a82a24afd4d4 (accessed 15/06/2012). UK Round 1 offshore wind farms. Proceedings of the
DNV (Det Norske Veritas) (2011) Design of Offshore Wind Institution of Civil Engineers – Energy 163(4): 167–181.
Turbine Structures. DNV, Oslo, Norway, DNV-OS-J101. Fischer T (2011) Executive Summary – UpWind Project. WP4:
Dobry R, Vincente E, O’Rourke M and Roesset J (1982) Offshore Foundations and Support Structures. See http://
Stiffness and damping of single piles. ASCE Journal of www.upwind.eu/pdf/WP4_Executive_Summary_Final.pdf
Geotechnical Engineering 108(3): 439–458. (accessed 12/09/2012).
Douglas-Westwood Limited (2002) The World Offshore Frandsen S and Thøgersen M (1999) Integrated fatigue
Renewable Energy Report 2002–2007. See http:// loading for wind turbines in wind farms by combining

150
Energy Offshore wind-turbine structures:
Volume 166 Issue EN4 a review
Arshad and O’Kelly

ambient turbulence and wakes. Wind Engineering 23(6): LeBlance C (2009) Design of Offshore Wind Turbine Support
327–339. Structures – Selected Topics in the Field of Geotechnical
Genta G (1998) Vibration of Structures and Machines, 3rd edn. Engineering. PhD thesis, Aalborg University, Denmark.
Springer, New York, NY, USA. Letchford C and Zachry B (2009) On wind, waves, and surface
GL (Germanischer Lloyd) (2005) Rules and Guidelines, IV drag. Proceedings of the 5th European and African Conference
Industrial Services, 2 Guideline for the Certification of on Wind Engineering, Florence, Italy, pp. 83–114.
Offshore Wind Turbines, 6 Structures. See http://onlinepubs. Little RL and Briaud JL (1988) Full Scale Cyclic Lateral Load
trb.org/onlinepubs/mb/Offshore%20Wind/Guideline.pdf Tests on Six Single Piles in Sand. Geotechnical Division,
(accessed 17/09/2012). Texas A&M University, College Station, TX, USA,
GWEC (Global Wind Energy Council) (2011) Global Wind miscellaneous paper GL–88–27.
Report – Annual Market Update. See http://www. Long J and Vanneste G (1994) Effects of cyclic lateral loads
windpower-international.com/microsites/gwec/Annual_ on piles in sand. ASCE Journal of Geotechnical Engineering
report_2011_lowres.pdf (accessed 14/09/2012). 120(1): 225–244.
Haritos N (2007) Introduction to the analysis and design of Malhotra S (2011) Design and construction considerations for
offshore structures – an overview. Electronic Journal of offshore wind turbine foundations in North America. In
Structural Engineering 7(Special Issue): 55–65. Wind Turbines (Al-Bahadly I (ed.)). InTech, Rijeka,
Henderson A, Morgan C, Smith B et al. (2003) Offshore wind Croatia, pp. 231–264.
energy in Europe – a review of the state-of-the-art. Wind Maria IB (2009) The economics of wind energy. Renewable
Energy 6(1): 35–52. and Sustainable Energy Reviews 13(6–7): 1372–1382.
IEA (International Energy Agency) (2009) Annual Report Martin J, André F and Wim CT (2004) Cost reduction prospects
2008. See http://www.ieawind.org/Annual_reports_PDF/ for offshore wind farms. Wind Engineering 28(1): 97–118.
2008/2008%20AR_small.pdf (accessed 10/09/2012). Matlock H (1970) Correlation for design of laterally loaded
IEA (2011) China Wind Energy Development Roadmap 2050. piles in soft clay. Proceedings of the Offshore Technology
See http://www.iea.org/publications/freepublications/ Conference, Houston, TX, USA, pp. 577–594.
publication/china_wind.pdf (accessed 01/03/2013). Musial W and Bonnie R (2010) Large-scale Offshore Wind
IEC (International Electrotechnical Commission) (2005) IEC Power in the United States: Assessment of Opportunities and
61400–1: Wind turbines – Part 1: Design requirements, Barriers. See http://www.nrel.gov/docs/fy10osti/49229.pdf
3rd edn. IEC, Geneva, Switzerland. (accessed 12/07/2012).
Igoe D, Gavin K and O’Kelly B (2013) An investigation into the Musial W and Butterfield S (2006) Energy from Offshore
use of push-in pile foundations by the offshore wind sector. Wind. See http://www.nrel.gov/docs/fy06osti/39450.pdf
International Journal of Environmental Studies, http://dx.doi. (accessed 15/07/2012).
org.10.1080/00207233.2013.798496 (accessed 01/10/2013). Quarton D, Rasmussen F, Argyriadis K and Nath C (1996)
IRENA (International Renewable Energy Agency) (2012) Wind turbine design calculations – the state of the art.
Renewable Energy Technologies: Cost Analysis Series, Proceedings of the European Wind Energy Conference,
http://www.irena.org/Document.Downloads/Publications/ Göteborg, Sweden, pp. 10–15.
RE_Technologies_Cost_Analysis-WIND_POWER.pdf Reese LC, Cox WR and Koop FD (1974) Analysis of laterally
(accessed 15/05/2012). loaded piles in sand. Proceedings of the 6th Annual Offshore
Ismael N (1990) Behaviour of laterally loaded bored piles in Technology Conference, Houston, TX, USA, pp. 473–484.
cemented sands. ASCE Journal of Geotechnical Engineering Reese LC, Cox WR and Koop FD (1975) Field testing and
116(11): 1678–1699. analysis of laterally loaded piles in stiff clay. Proceedings of
Journée JMJ and Massie WW (2001) Offshore Hydrodynamic, the 7th Annual Offshore Technology Conference, Dallas, TX,
1st edn. Delft University of Technology, Delft, the USA, pp. 671–690.
Netherlands. Roddy DJ, Yu Y, Dufton DJ and Thornley P (2009) Low-
Junginger M, Agterbosch S, Faaij A and Turkenburg W carbon energy solutions for an ecological island in China.
(2004) Renewable electricity in the Netherlands. Energy Proceedings of the Institution of Civil Engineers – Energy
Policy 32(9): 1053–1073. 162(2): 85–95.
Kooijman H, Noord MD, Volkers C et al. (2001) Cost and Rodenhausen M (2010) Soil Response of Offshore Wind
potential of offshore wind energy on the Dutch part of the Turbines. MSc thesis, University of Stuttgart, Stuttgart,
North Sea. Proceedings of the European Wind Energy Germany.
Conference and Exhibition, Copenhagen, Denmark. Saleem Z (2011) Alternatives and Modifications of Monopile
Krohn S, Morthrost P and Awerbuch S (eds) (2009) Foundation or its Installation Technique for Noise
Economics of Wind Energy. European Wind Energy Mitigation. See www.vliz.be/imisdocs/publications/223688.
Association, Brussels, Belgium. pdf (accessed 03/08/2012).

151
Energy Offshore wind-turbine structures:
Volume 166 Issue EN4 a review
Arshad and O’Kelly

Schaumann P and Böker C (2005) Can Tripods and Jackets Treehugger (2011) Is Direct Drive the Future? Wind Turbines
Compete with Monopiles? See http://wind.nrel.gov/public/ Without Gears Are Lighter, Cheaper, More Reliable. See
SeaCon/Proceedings/Copenhagen.Offshore.Wind.2005/ http://www.treehugger.com/renewable-energy/is-direct-
documents/papers/Low_cost_foundations/P.Schaumann_ drive-the-future-wind-turbines-without-gears-are-lighter-
Can_jackets_and_tripods_compete_with_monopile.pdf cheaper-more-reliable.html (accessed 15/11/2012).
(accessed 06/08/2012). Tricklebank AH (2008) Briefing: Offshore wind energy – a
Søren K, Poul-Erik M and Shimon A (2009) The Economics of challenge for UK civil engineering. Proceedings of the
Wind Energy. See http://www.windenergie.nl/sites/ Institution of Civil Engineers – Energy 161(1): 3–6.
windenergie.nl/files/documents/the_economics_of_ Vries WE (2007) UpWindProject.WP4: Assessment of
windenergy_ewea.pdf (accessed 17/07/2012). Bottom-mounted Support Structure Types with Conventional
Sørensen SPH, Brødbæk KT, Møller M, Augustesen AH and Design Stiffness and Installation Techniques for Typical
Ibsen LB (2009) Evaluation of load–displacement Deep Water Sites. See http://www.upwind.eu/pdf/Upwind_
relationships for large-diameter piles in sand. In WP4_D4.2.1_%20Assessment%20of%20bottom-
Proceedings of the 12th International Conference on Civil, mounted%20support%20structure%20types.pdf (accessed
Structural and Environmental Engineering Computing 12/05/2012).
(Topping BHV, Costa Neves LF and Barros RC (eds)). Vries WE (2011) UpWindProject.WP4: Support Structure
Civil-Comp Press, Sterling, UK, paper 244. Concepts for Deep Water Sites. See http://www.upwind.eu/
Tarp-Johansen NJ (2005) Partial safety factors and pdf/Final%20report%20WP4.2.pdf (accessed 10/08/2012).
characteristic values for combined extreme wind and wave Wiser R and Bolinger M (2011) 2010 Wind Technologies
load effects. Journal of Solar Energy Engineering 127(2): Market Report. US DOE, Office of Energy Efficiency and
242–252. Renewable Energy, Washington, DC, USA.
Tempel JV (2006) Design of Support Structures for Offshore WWEA (World Wind Energy Association) (2011) World Wind
Wind Turbines. PhD thesis, Delft University of Technology, Energy Report 2010. See http://www.wwindea.org/home/
Delft, the Netherlands. images/stories/pdfs/worldwindenergyreport2010_s.pdf
Tempel JV and Molenaar DP (2002) Wind turbine structural (accessed 10/08/2012).
dynamics – a review of the principles for modern power YFH (Yuanta Financial Holdings) (2011) Industry Update.
generation, onshore and offshore. Wind Engineering 26(2): Greater China: Energy. See http://ipreo.YUANTA.com/
211–220. NSightWeb_v2.00/Downloads/Files/29907.pdf (accessed
Tong W (2010) Wind Power Generation and Wind Turbine 17/07/2012).
Design. WIT Press, Southampton, UK.

WHAT DO YOU THINK?


To discuss this paper, please email up to 500 words to the
editor at journals@ice.org.uk. Your contribution will be
forwarded to the author(s) for a reply and, if considered
appropriate by the editorial panel, will be published as a
discussion in a future issue of the journal.
Proceedings journals rely entirely on contributions sent in
by civil engineering professionals, academics and students.
Papers should be 2000–5000 words long (briefing papers
should be 1000–2000 words long), with adequate illus-
trations and references. You can submit your paper online
via www.icevirtuallibrary.com/content/journals, where you
will also find detailed author guidelines.

152
International Journal of Physical Modelling International Journal of Physical Modelling in Geotechnics,
2014, 14(3), 54–67
in Geotechnics http://dx.doi.org/10.1680/ijpmg.13.00015
Volume 14 Issue 3 Paper 1300015
Received 21/11/2013 Accepted 28/05/2014
Development of a rig to study model pile Keywords: foundations/models (physical)/wind loading & aerodynamics
behaviour under repeating lateral loads
Arshad and O’Kelly
ICE Publishing: All rights reserved

Development of a rig to study model


pile behaviour under repeating
lateral loads
&1 Muhammad Arshad MSc, ME &2 Brendan C. O’Kelly PhD, FTCD, CEng, CEnv, MICE
PhD Candidate, Department of Civil, Structural and Environmental Associate Professor, Department of Civil, Structural and Environmental
Engineering, Trinity College Dublin, Dublin, Ireland Engineering, Trinity College Dublin, Dublin, Ireland

1 2

This paper deals with an important problem of the effect of repeating lateral loads on large-diameter monopile
foundations for offshore wind turbine (OWT) structures. The cycles of typically low-amplitude repeating lateral
loads and moments generated by various environmental factors are resisted by horizontal earth pressures
mobilised in the soil surrounding the pile. Laboratory testing on reduced-scale piles is an efficient and economical
way to investigate such pile–soil behaviour. This paper presents the development of a new mechanical loading
system to apply many thousands of repeating cycles of lateral load in different forms to a 1g model, with full
control provided over the loading direction (i.e. one-way or two-way lateral loading), amplitude, frequency and
waveform shape (e.g. sinusoidal, square or haversine). Compared with equivalent setups employing pneumatic or
hydraulic actuators, the new loading system is able to produce similar performance at lower cost and also provides
more control over the waveform shape. A programme of lateral load tests, each involving many thousands of load
cycles, was performed on a rigid model pile installed in dry sand beds to demonstrate some of the main capabilities
of the new system.
^
Notation h non-dimensional parameter for pile rota-
D pile outer diameter tion from its initial alignment
Dl spring extension
d50 mean particle size
^
e non-dimensional parameter for moment
arm 1. Introduction
EI pile bending stiffness At present, there is some uncertainty in the literature regarding
g gravitational constant the change in stiffness of the soil surrounding a monopile under
H applied lateral load long-term repeated lateral loading. In the case of rigid
^
H non-dimensional parameter for lateral load monopiles, the soil would fail rather than the piles failing by
L pile embedment length plastic hinges. For sandy soil, the API (2010) and DNV (2011)
M bending moment applied at pilehead models always predict degradation of the absolute secant
^
M non-dimensional parameter for bending stiffness under repeated lateral loading, irrespective of the
moment density state or number of load cycles. Achmus et al. (2009)
Nh vertical distance between loading node reported degradation of the absolute secant stiffness for medium
position and the bottom of chain loop dense and dense sands with increasing number of load cycles.
when sliding node is at its mean position However model studies on monopiles in loose (Bhattacharya
pa atmospheric pressure et al., 2011; LeBlanc et al., 2010) and dense (Cuéllar et al., 2012,
c9 effective unit weight Rosquoet et al., 2007) sands suggested that the foundation

54
International Journal of Physical Modelling in Geotechnics Development of a rig to study
Volume 14 Issue 3 model pile behaviour under
repeating lateral loads
Arshad and O’Kelly

stiffness (cyclic) increases with the number of load cycles. the lateral pile deflection at seabed level to 0?11 m (Malhotra,
LeBlanc (2009), Bhattacharya and Adhikari (2011) and Cuéllar 2011) or the pile rotation from its initial (vertical) alignment to
et al. (2012) suggested that the stiffness increase occurs as a 0?5 ˚ (LeBlanc, 2009).
result of densification of the soil next to the pile. There are also
differences of opinion regarding the rate of accumulation of The foundation response under repeated lateral loading is a
lateral strain in the surrounding soil for piles subjected to one- major design consideration. An economical design procedure
way or two-way lateral loading. Under one-way loading, the can be efficiently and confidently applied if experimental
magnitude of the load wave fluctuates between zero and the verifications are achievable, either through full-scale testing or
applied maximum load amplitude. In response, the pile tends to model studies. In situ full-scale pile tests are expensive and
deflect (rotate) only in one direction from its initial (vertical) time-consuming to perform. Hence the literature contains
alignment. Under two-way loading, the pilehead is sequentially many reduced-scale model studies performed in the field (e.g.
pulled in opposite directions, such that the pile tends to deflect Gavin and O’Kelly (2007) and Igoe et al. (2013)) or in the
(rotate) alternately in opposing directions from its mean vertical laboratory employing different types of loading systems to
position (Peng et al., 2011). For two-way lateral loading, the validate numerical predictions and design rules, such as given
accumulative tilt of the pile usually follows the direction of the by Achmus et al. (2009).
first quarter of the loading cycle (Long and Vanneste, 1994;
Rosquoet et al., 2007). Some researchers (LeBlanc et al., 2010;
Model pile tests involving static lateral loading have been
Klinkvort and Hededal, 2013; Zhu et al., 2013) have shown that
performed to determine ultimate lateral capacity (e.g. Reese
two-way lateral loading produces higher rates of accumulation
et al. (1974) and Ramakrishna and Rao (1999)) and to
of lateral strain in the surrounding soil compared with one-way
investigate the influence of the piles’ cross-section and rigidity
lateral loading. Long and Vanneste (1994), Lin and Liao (1999)
on lateral load-carrying capacity (Mahmoud and Burley, 1994;
and Peng et al. (2011) were of the opinion that the opposite was
Rao et al., 1998; Raongjant and Meng, 2011). Some of these
the case. The findings of such studies awere based on results
studies have focused on investigating the behaviour of tapered
generated by experimental loading systems having limited
piles (Dührkop et al., 2010; El Nagger and Wei, 1999), finned
operational capabilities, with all of these studies considering
piles (Peng et al., 2011) and (or) piles subjected to many
only one waveform shape (i.e. sinusoidal) and for a particular
loading frequency. thousands of load cycles in the lateral direction (Dührkop et al.,
2010; LeBlanc et al., 2010; Peng et al., 2011; Cuéllar et al., 2012).
Model testing results can be directly applied once the soil at site
In reality, the pile deflection (rotation) response under lateral
is the same for the scaled tests. The best way is to link the model
loading arises from the soil behaviour and also from dynamic
tests with element test parameters. For example, Lombardi et al.
pile–soil interaction, which are dependent on the loading history
(2013) linked their model tests with threshold strains that can be
and applied loading conditions. There are many examples of
obtained from standard element tests such as the resonant
structures supported on pile foundation systems for which
lateral loading may prove to be critical over gravitational forces. column.
For instance, in a study related to offshore wind turbines
(OWTs) supported by monopile foundation systems in the The reliability (acceptability) of laboratory model studies
North Sea, it was found that the magnitude of dynamic lateral depends, among other factors, on the accuracy and efficiency
loading can be up to 66% of the gravitational loads (Byrne and of the loading system in simulating the actual field loading
Houlsby, 2003). This scenario is more onerous when the lateral conditions through certain scaling laws. In the case of model
loading is continuously repeating at varying frequency, load studies for offshore structures, the experimental loading system
amplitude and direction (Arshad and O’Kelly, 2013). At some must be capable of simulating the environmental loading,
critical level of load amplitude and/or frequency, these repeating particularly that acting in the horizontal direction, with
lateral loads can cause significant reductions in the lateral soil varying amplitude and frequency.
resistance for a monopile foundation structure (Ramakrishna
and Rao, 1999). The required lateral load-carrying capacity of This paper describes the development of a novel mechanical
the pile depends on the type and purpose of the structure in loading system for the application of many thousands of lateral
which it is an integral part. The rotation of the pile from its loading cycles, with full control provided over the direction,
initial (vertical) alignment must be limited in the range 0?17–2?0 ˚ amplitude, frequency and waveform shape. Experimental
(Lee et al., 2010, 2011) for proper functioning of different types data for repeated lateral loading of an instrumented model
of supported structures. pile installed in dry sand are presented to demonstrate the main
capabilities of the new loading system. The effects of different
Under field loading, the proper operation of OWTs supported loading scenarios (load directions, amplitudes and frequencies)
by monopile foundation systems generally necessitates limiting are also investigated.

55
International Journal of Physical Modelling in Geotechnics Development of a rig to study
Volume 14 Issue 3 model pile behaviour under
repeating lateral loads
Arshad and O’Kelly

2. Forcing frequency, fundamental natural the region after the 3P range as the ‘stiff–stiff’ zone. ‘Soft–soft’
frequency and resonance and ‘stiff–stiff’ zones are unsuitable for the design solution.
It is essential to consider the fundamental natural frequency of a The structure is considered too flexible if its natural frequency
wind-turbine structure for a proper description/evaluation of its falls within the ‘soft–soft’ zone and too rigid (heavy and
dynamic behaviour. In the literature related to OWTs, the term expensive) within the ‘stiff–stiff’ zone. Another important
‘cyclic loading’ is generally used in connection with the reason for avoiding the ‘soft–soft’ frequency region is that
harmonic variation of the repeating load wave in time domain. wind turbulence and wave excitation frequencies usually fall
When the forcing (loading) frequency gets closer to the within this zone (LeBlanc, 2009). For wind-turbine structures,
structure’s natural frequency, the repeating load can be termed this invariably leads to the development of higher stresses in
as dynamic load, which tends to excite the structure dynami- the support structure/foundation, which is an unfavourable
cally, leading to resonance. Dynamic effects can be considered situation in considering fatigue life. This dynamic interaction of
simplistically in a linear system through dynamic amplification the soil–pile system produces greater strain in the surrounding
factors. However this approach is strictly not valid for an OWT soil and may lead to higher stiffness degradation (Achmus et al.,
foundation structure owing to various non-linear ties in the 2009) or soil stiffening (Cuéllar et al., 2012; LeBlanc et al., 2010).
system, as shown experimentally by Bhattacharya et al. (2012), It is therefore important to ensure that excitation frequencies
although the error may not be great. Hence, engineers need to having high energy levels do not coincide with the support
adjust the cyclic load results to take into account the dynamic structure’s natural frequency. DNV (2011) suggests that the
behaviour of the system. natural frequency of the structure should not come close to the
1P or 3P frequency regions, remaining away from these zones by
Major sources of forcing frequencies for OWTs are wind, a margin of at least 10%. A detailed insight into the dynamic of
wave, current and any out-of-line imbalances of the rotating the soil-structure behaviour for OWTs is given by Adhikari and
electromechanical parts installed in the turbine’s nacelle-rotor Bhattacharya (2011, 2012), Bhattacharya et al. (2011, 2012,
system (Bhattacharya et al., 2011). For a typical 5 MW OWT 2013), Bhattacharya and Adhikari (2011) and Lombardi et al.
having rotational speeds of typically 10–20 rpm for the main (2013).
rotor shaft, the first excitation frequency 1P (i.e. corresponding
to one full revolution) occurs in the range 0?17–0?33 Hz. For a 3. Scaling issues for model testing
three-bladed turbine, the blade passing frequency of typically To relate the results from reduced-scale physical modelling
0?5–1?0 Hz is denoted as the 3P frequency, which is heavily with the corresponding results of the prototype system, certain
excited on account of the impulse-like excitation arising from conditions must be met to preserve constitutive and kinematic
the blades passing by the tower. In the North Sea environment, similarities between the systems. Examples of the derivation of
excitation from wind and waves usually occurs with frequen- scaling laws for offshore monopile foundations have been
cies of 0?01 and 0?1 Hz, respectively, as illustrated in Figure 1. reported by Lai (1989), Muir Wood et al. (2002), LeBlanc et al.
(2010), Bhattacharya et al. (2011) and Cuéllar et al. (2012).
The region of ‘wanted frequency’ (i.e. between the 1P and 3P However, the satisfaction of such scaling conditions for
ranges) identified in this figure is referred to as the ‘soft–stiff’ similarity in granular soils is not trivial (Bhattacharya et al.,
zone; the region before the 1P range as the ‘soft–soft’ zone and 2011; Dong et al., 2001), especially for a complex problem such
as OWTs where one has dynamics, aerodynamics, soil–
structure interaction and fluid flow around the foundations.
Wind turbulance Waves

Among other technical difficulties of performing physical tests


with geomaterials, the direct scaling of the particle grain
Power spectral density

dimensions is particularly problematic as it may introduce


1P Wanted 3P undesired forces into play unless a certain minimum ratio is
frequency
maintained between mean grain size (d50) and a characteristic
dimension (pile diameter in the present investigation) of the
model (Sedran et al., 2001; Verdure et al., 2003). Moreover,
constitutive laws describing the load–deformation behaviour of
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 the soil are stress-level dependent, which implies that for 1g
Frequency: Hz conditions, homologous points of the soil foundations for the
model and prototype may experience different deformational
Figure 1. Typical excitation frequency ranges for 5 MW offshore responses. For instance, the load–deformation response of
wind-turbine structure structures founded on sand is governed, among other factors,
by the magnitude of mean effective confining pressure, which is

56
International Journal of Physical Modelling in Geotechnics Development of a rig to study
Volume 14 Issue 3 model pile behaviour under
repeating lateral loads
Arshad and O’Kelly

different for the model and prototype. Laboratory model tests The novel rig for model pile studies developed at Trinity
on sand at 1g (i.e. at low confining pressure) are generally College Dublin (TCD) which is presented in this paper is not
performed at a lower density index compared with full-scale in only capable of applying many thousands of load cycles under
order to match the angle of shearing resistance and take into one-way and two-way loading conditions but is also efficient in
account the dilation (LeBlanc et al., 2010). Similitude relation- generating different waveform shapes within the frequency
ships in the form of non-dimensional parameters (NDPs) can be range compatible with offshore environmental loading.
applied to relate the results for the model with the prototype,
and vice versa (see LeBlanc et al., 2010). These NDPs consider 5. Development of the new rig for model
the pile’s geometry, material properties, applied loading and its pile studies
lateral displacement (rotation) response, as given in Table 1,
with due consideration of the pile tip displacement under 5.1 General layout and components
repeating lateral loads. Figure 2 shows the general layout of the new rig that is used to
apply repeating lateral loads to the head of an instrumented model
pile. The pile (identified as (15) in the figure) is installed in a soil
4. Existing systems for lateral loading of
bed that is contained in a steel tank (8) having an internal diameter
model piles
of 0?95 m, overall depth of 0?60 m, and which is reinforced by box
Over the past two decades, a number of different systems have
sections welded around its outer perimeter. The driving torque to
been developed to apply monotonic and (or) repeating lateral the loading system is provided by a three-phase 1?5 kW AC motor
loads in performing laboratory studies on model piles. The (1), which torques gearing through a system of metallic chain
methods of operation of these systems can be broadly linkages – that is, a bicycle chain. The drive motor and gearing are
categorised as mechanical (Rao et al., 1998), electromechanical mounted on a reaction frame (9). The steel tank and soil bed
(Basack, 2005; LeBlanc, 2009; LeBlanc et al., 2010), pneumatic within are isolated from the reaction frame and hence any
and (or) hydraulic (Chandrasekaran et al., 2010; El Nagger and vibrations produced by the operation of the mechanical loading
Wei, 1999; Kumar and Rao, 2012). As discussed and presented system. In the present set-up, the chains had a pitch length of
in these experimental studies, the loading systems have varying 12?7 mm, pin length of 6?0 mm and roller diameter of 7?2 mm.
complexities in their operation and capabilities to simulate The speed control unit for the motor provides control over the
the field conditions. For instance, some loading systems (Rao speed of rotation of the motor shaft and hence the loading
et al., 1998; El Nagger and Wei, 1999) can apply only frequency. Rotation of gear x (3) moves the sliding node (13) to-
monotonic lateral loading in a certain range. Others can apply and-fro between points L and R. The sliding node is connected to
one-way repeating load and have been used in studies to apply gear x by an articulated arm (12). The travel range of the sliding
limited numbers of lateral loading cycles (Chandrasekaran node is controlled by adjusting the location of the pivot point (pin
et al., 2010; Kumar and Rao, 2012; Qin, 2010). A few systems joint) and the arm length. Referring to the right-hand side (RHS)
(Basack, 2005; LeBlanc et al., 2010; Peng et al., 2006) have of the schematic arrangement shown in Figure 2: one end of the
been designed to apply many thousands of load cycles for one- RHS chain is connected to the sliding node and its other end to the
way and two-way loading conditions, with sufficient control RHS of the pilehead by way of a miniature load cell (10) and
provided over the load amplitude and frequency. A detailed spring (11). As shown in the figure, this chain loops, engaging with
discussion of existing loading systems is beyond the scope of 80 mm diameter upper (2) and lower (5) cogs, from which a dead
this paper and the reader is referred to the publications cited weight (7) can be attached to a specific chain linkage by way of a
above for further information. load hanger. Under two-way loading, a similar arrangement is
used to apply and measure the loading to the left-hand side (LHS)
of the pilehead. The loading system can be operated from a very
Moment loading ^ M
M~ low frequency of 0?01 Hz up to 2?0 Hz by simply adjusting the
L3 Dc’
speed of rotation of the drive motor. Another determining factor
Horizontal loading ^ H
H~ 2 in achieving the desired frequency is the stiffness of the spring(s).
L Dc’
rffiffiffiffiffiffi Loading frequencies in the range 0?1–2?0 Hz have been used by
Rotation ^ pa
h~ different researchers (Cuéllar et al., 2012; LeBlanc et al., 2010;
Lc’
rffiffiffiffiffiffi Lombardi et al., 2013; Peng et al., 2011) in performing model
Moment arm (i.e. vertical distance between M
^
e~ studies at 1g on monopile foundations for OWTs.
initial level of sand bed surface and point of HL
horizontal load application)
5.2 Working mechanism
Table 1. Non-dimensional parameters for scaling laws (after The working mechanism of the new loading system depends on
LeBlanc et al., 2010) the required loading scheme, with the range of possibilities
described in the next section of the paper. Referring to

57
International Journal of Physical Modelling in Geotechnics Development of a rig to study
Volume 14 Issue 3 model pile behaviour under
repeating lateral loads
Arshad and O’Kelly

L O R

2
3
4

5
Sand level in steel tank
10
11 6
12

13

7
14
15
8

Figure 2. Schematic diagram of the TCD rig for model pile studies, hanger; 8, steel tank; 9, reaction frame; 10, right load cell; 11, right
set up for two-way lateral loading. Key: 1, drive motor; 2, upper spring; 12, articulated arm; 13, sliding node; 14, reference support
right cog; 3, gear x; 4, upper right chain segment; 5, lower right with LVDTs; 15, pile
cog; 6, lower right chain segment; 7, right loading weight with

Figure 3, repeating lateral loading behaviour is obtained position for the RHS chain loop located on the upper chain
through the transfer of loads (supplied by the loading weights segment (4), the tension force applied by the loading weight
and hangers) from the upper chain segments to the lower chain and hanger (7) remains in the upper RHS chain segment for
segments, and vice versa; that is, tension forces are generated in a greater proportion of the load cycle. When the sliding
the upper and lower chains alternately. The RHS lower chain node (2) moves to location R, the bottom section of the
segment (6) goes from the spring (8) to the bottom of the chain RHS chain loop is at its lowest position.
loop, passing over the lower cog (5). The RHS upper chain (b) The stiffness of the connecting spring(s) (8) which contributes
segment (4) goes from the sliding node (2) to the bottom of the to generating the required waveform shape. If the spring(s)
chain loop, passing over the upper cog (3). is removed (i.e. lower chain segment(s) connected directly to
the pilehead), then an almost square wave form is produced
In addition to the applied hanger loads and speed of the drive by the loading system owing to the abrupt action and
motor, other governing inputs for the loading scheme are as interaction of the loading to the pilehead. In other words, for
outlined below. a particular load amplitude, the loading frequency is also
determined by the response time (stiffness) of the spring(s).
(a) The locations of the attachment points for the RHS and
(or) LHS load hangers along the chain loops, with reference A synchronised adjustment/correlation between the range of
to the centres of the upper cogs; that is, the loading node node movement (i.e. points L to R in Figure 3) and elongated
position on the vertical chain-loop length. Referring to length(s) of the spring(s) under loading to get the desired
Figure 3, the vertical chain-loop length (VCLL) equals the loading scheme is a difficult task. This is more onerous for two-
horizontal chain-loop length (HCLL) when the sliding node way than one-way loading. This task was completed incremen-
(2) is at its mid-range (i.e. point O), which represents zero tally by experimental trials.
wave amplitude. The loading node position determines the
loading impact on the pilehead; that is, the duration and For demonstration purposes, Tables 2 and 3 show the relative
portion of the load cycle for which a particular chain positions of the sliding node and loading hangers at interim
segment (upper or lower) remains under the action of the stages for loading under one-way and balanced two-way
applied tension force. For example, with the loading node sinusoidal load waves.

58
International Journal of Physical Modelling in Geotechnics Development of a rig to study
Volume 14 Issue 3 model pile behaviour under
repeating lateral loads
Arshad and O’Kelly

2 HCLL

L O R

VCLL
9 8 6
7

10 Loading node position

Nh
Bottom of chain loop
when sliding node is at
midrange (point O)

Figure 3. Schematic diagram of new loading system, set up for right cog; 6, lower right chain segment; 7, right loading hanger
two-way lateral loading. Key: 1, slot for sliding node; 2, sliding with weights; 8, right spring; 9, right load cell; 10, pilehead; HCLL
node; 3, upper right cog; 4, upper right chain segment; 5, lower and VCLL, horizontal and vertical chain-loop lengths, respectively

Relative position
Sliding node Spring of chain and Load cell
Stage position response load hanger measurement Remark

1 L O R Dl Load cycle starts from maximum applied load


amplitude, with lower right chain segment under
Load

Time
tension.

2 L O R Extension in spring (and hence tension force)


reduces as sliding node moves towards the LHS.
Load

Time

3 L O R Dl = 0 With the sliding node at point L, the spring has


returned to its original length and the hanger load
Load

Time
tends to shift onto the upper right chain segment.

4 L O R As the sliding node moves from points L to R, the


spring extends and the load applied to the
Load

Time
pilehead increases again.

5 L O R When the sliding node reaches point R, the spring


is under maximum elongation and the maximum
Load

Time
lateral load is applied to the pilehead, completing
one full sinsuoidal cycle.

Table 2. Reference positions of moving parts in loading system for


one-way sinusoidal wave

59
60
Volume 14 Issue 3

Sliding Response Response Position of Position of Response Response


node of LHS of RHS LHS chain and RHS chain and of LHS of RHS
Stage position spring spring load hanger load hanger load cell load cell Remark

1 L O R Dl = 0 Dl At point R, the RHS spring is at its full elongation, with the full
weight of the RHS hanger acting on the RHS lower chain segment.
Time

Load

Load
Time

2 L O R At point O, the RHS spring has returned to its original length.


Further movement of the sliding node towards point L causes

Load

Load
Time
International Journal of Physical Modelling in Geotechnics

the LHS spring to extend.

3 L O R Dl At point L, the LHS spring is at its full elongation, with the full
weight of the LHS hanger acting on the LHS lower chain

Load
Load
Time
segment.

4 L O R At point O, the LHS spring has returned to its original length.


Further movement of the sliding node towards point R causes

Load
Load
Time
the RHS spring to extend.
Arshad and O’Kelly

5 L O R At point R, the RHS spring is at maximum elongation under


the lateral force equivalent to that of the RHS load hanger and
repeating lateral loads

Load

Load
Time
weights.
model pile behaviour under
Development of a rig to study

Table 3. Reference positions of moving parts to produce sinusoidal


wave under balanced two-way loading
International Journal of Physical Modelling in Geotechnics Development of a rig to study
Volume 14 Issue 3 model pile behaviour under
repeating lateral loads
Arshad and O’Kelly

5.2.1 One-way sinusoidal load wave segment applies zero force to the pilehead. As the sliding node
For this configuration, only one load hanger with dead weights is moves within the left-half span, the tension force in the LHS
attached to the chain system (RHS chain in the scenario lower chain segment increases. When the sliding node reaches
considered below (see Table 2 and Figure 2)). During set up, point L, the pilehead experiences the maximum load amplitude
with the sliding node positioned at point L, the RHS load hanger acting towards the LHS (stage 3, Table 3). A full sinusoidal
is attached to the chain link at the bottom of the RHS chain loop cycle is completed when the sliding node reaches point R again;
system. The range of sliding node movement (i.e. distance that is, the maximum load amplitude is achieved in the RHS
between points L and R) is adjusted such that it equals the lower chain segment (stage 5, Table 3).
extension (Dl) of the RHS spring, located between the lower chain
segment and load cell, in response to the applied load. For 5.3 Capabilities of the new loading system
instance, suppose that at the start of the load cycle considered, the The new loading system was designed to generate many
sliding node is at its extreme right position (point R), in which thousands of load cycles of varying load amplitude, frequency
case the RHS spring is elongated by length Dl under the force and waveform shape. The load amplitude depends on the
applied by the RHS load hanger with weights (see stage 1, magnitudes of the weights applied to the RHS and (or) LHS
Table 2). At this stage, the pilehead experiences the maximum load hangers. Referring to Figure 3, the required waveform
lateral tension force. With the movement of the sliding node from shape (i.e. sinusoidal, square, haversine etc.) can be obtained
point R towards point L (stage 2, Table 2), the extension in the by adjusting the length Nh (i.e. distance between the loading
spring reduces, such that when the sliding node is positioned at its node position and the bottom of the chain loop when the
extreme left position (point L), the spring experiences no sliding node is at midrange) and (or) by using springs with
extension (stage 3, Table 2); in other words, no lateral load is different stiffnesses. Control over the frequency of the load
applied to the pilehead. At this moment, the tension force in the wave is primarily achieved by adjusting the speed of the drive
upper right chain segment due to the load hanger with weights is motor and to a less extent by the spring stiffness. The loading
completely carried by the torque of the drive motor. With the scheme can be arranged to produce
sliding node moving from point L to R, the spring extends and the
lateral load applied to the pilehead increases again, reaching its (a) one-way loading, achieved when only one load hanger
maximum value when the sliding node reaches point R (stage 5, with dead weights is used, with the magnitude of the load
Table 2), thereby completing one full sinusoidal cycle. For one- wave fluctuating between zero and the applied maximum
way sinusoidal wave loading, the vertical distance between the load amplitude
loading node position and the bottom of the chain loop remains (b) one-way loading, with the magnitude of the load wave
practically zero when the sliding node is at its mean position. fluctuating between the applied maximum load amplitude
and some value above the zero value, hereafter termed as
5.2.2 Balanced two-way sinusoidal load wave partial one-way loading
For this configuration, load hangers with weights are attached (c) balanced two-way loading when equal weights are applied
to the RHS and LHS chains (see Table 3 and Figure 3). During to the two load hangers
set-up, with the sliding node located at point O, the load hangers (d) unbalanced two-way loading when unequal weights are
with weights are attached at the bottom of the RHS and LHS applied to the two hangers.
chain loop systems. As long as the sliding node remains in the
right-half span of node movement (i.e. between points O and R), For balanced or unbalanced two-way loading of any waveform
the RHS lower chain segment and spring remain under varying shape, the RHS load cell shows zero force when the LHS load
tension force, with the pile tending to deflect towards the right cell measures a tension force, and vice versa. The resultant shape
side from its initial (mean) vertical position. Similarly, when the of the load wave for the whole system is obtained by
sliding node moves within the left-half span, a varying tension superimposing the load–time curves measured by both load
force acts in the LHS lower chain segment and spring, with the cells. For demonstration purposes, different shapes of repeating
pile tending to deflect towards the left side. During set-up, the lateral load waves were generated in the present study for
half-span length is adjusted such that it equals the extension (Dl) maximum load amplitudes in the range of 30–45% of the pile’s
of the spring at the maximum load amplitude; that is, under the ultimate lateral load-carrying capacity under monotonic loading.
action of lateral force equivalent to load hanger with weights. These percentages correspond to the fatigue and serviceability
With the sliding node located at point R, the maximum tension limit states respectively (DNV, 2011). The experimental scenarios
force acts in the RHS lower chain segment (see stage 1, Table 3). considered in this paper are presented qualitatively in Figure 5.
As the sliding node moves away from point R, the lateral tension
force applied to the RHS of the pilehead reduces, with the force The loading system can also be configured to generate two
shifting to the RHS upper chain segment. On reaching point O different types of load wave under two-way loading; for
(i.e. at midrange: stage 2, Table 3), the RHS lower chain example, the tension force to the LHS of the pilehead could be

61
International Journal of Physical Modelling in Geotechnics Development of a rig to study
Volume 14 Issue 3 model pile behaviour under
repeating lateral loads
Arshad and O’Kelly

applied as a sinusoidal wave and the load to its RHS could be dry density of 1577¡6 kg/m3 for the 15 sand beds prepared in the
applied as a haversine wave. Such scenarios can be generated present investigation (density index range of 70–74%). At
by using springs having different elastic moduli and (or) maximum density, the dry sand had a peak friction angle of 39˚,
changing the locations of the attachment points for the load determined from 60 mm square shearbox tests. Geometrically the
hangers on the chains. pile set-up, with a pile embedment length to outer diameter (L/D)
ratio of 6?8, is categorised as a short rigid pile, which encompasses
6. Experimental demonstrations L/D ratios of up to 10 (Peng et al., 2011; Tomlinson, 2001). This
An extensive programme of testing was performed on a model scenario typically represents a field monopile (made of steel) for an
pile installed in dry, dense sand beds to demonstrate the OWT foundation system at 1/100 scale. Regarding possible
performance and repeatability of the new loading system in boundary effects associated with the relative dimensions of the
producing one-way and two-way sinusoidal lateral loading tank and model pile under the repeated lateral loading scenarios
conditions. These tests were performed in a constant tempera- investigated in the present study: (a) a soil cushion (sand in our
ture environment at 20¡2 ˚C. case) having a depth of three to four times the pile diameter located
below the pile tip is considered sufficient to absorb the vertical
6.1 Model pile stress field (LeBlanc et al., 2010); (b) with the ratio of the tank
The 540 mm long model pile was manufactured from brass tubing diameter to pile diameter greater than 17, side wall boundary
having outer and inner diameters of 53?0 and 51?4 mm, effects were not significant (Davie and Sutherland, 1978; Rao et al.,
respectively, which produced a bending stiffness (EI) of 1996).
4?33 kN m2. Its lower end was closed using 3 mm thick brass
plate to represent a fully plugged tubular pile. The model pile was 6.3 Instrumentation and data acquisition
instrumented over its embedment length using eight strain gauges Figure 5 illustrates the arrangement of load cells and linear-
of type TML-PL-10-11, manufactured by Tokyo Sokki Kenkyujo variable displacement transducers (LVDTs) used to measure the
Co. Ltd, Japan. These gauges were arranged in a line at 50 mm load–displacement (rotation) behaviour of the model pile. The two
centre-to-centre spacings and attached to the pile’s outer wall horizontally mounted miniature load cells (series LCM-703, range
surface using an epoxy adhesive, with the lowermost gauge located of ¡250 N, manufactured by Omega Engineering Ltd, UK)
at a distance of 20 mm above the pile tip. They had gauge length recorded the magnitudes of the repeating lateral loads applied to
and width dimensions of 10 and 5 mm, respectively, a gauge factor the pilehead by the loading system. Along the section of the pile
of 2?07 and full Wheatstone Bridge circuit configuration. A thin protruding above the sand bed surface, two horizontally mounted
cover layer of epoxy was applied over the attached gauges, and also LVDTs, one located 50 mm directly above the other, recorded
locally on their lead connections, in order to provide protection lateral displacements of the pilehead (see Figure 4). This arrange-
from potential damage during handling of the pile, its installation ment allowed the determination of the rotation (tilt) of the rigid pile
in the sand beds and under repeated lateral loading of the pile. from its initial vertical alignment, the depth to its point of rotation
along its embeded length and its lateral displacement at the sand
6.2 Sand characterisation and sand bed preparation, bed surface level. A third LVDT, mounted coaxially with the pile,
including pile installation measured the pile’s vertical displacement response. The LVDTs
In the present investigation, the model pile was partially embedded (series TR-0050, manufactured by Novotechnik Ostfildern,
in dense sand beds contained within the 0?95 m diameter by 0?6 m Germany) had a maximum range of 50 mm with linearity up to
deep steel tank. The sand beds were prepared using dry sub- 0.075%. Figure 6 shows photographs of the partially embedded
angular to angular medium silica sand having a d50 of 0?27 mm, instrumented pile undergoing two-way lateral loading applied by
coefficient of uniformity of 1?85 and coefficient of curvature of 1?0. the newly developed apparatus.
The sand was air-pluviated into the tank, raining in six layers, each
100 kg in mass, which produced deposited layers of approximately The outputs from the load cells, LVDTs and strain gauges were
90 mm in thickness. After depositing the first two layers, the model recorded by a System-7000 data acquisition system (Vishay
pile was aligned vertically at the centre of the tank, with temporary Precision Group, USA). This system simultaneously scanned
support to the pilehead provided by tensioned horizontal steel each sensor/channel at ten data points/s, with a measurement
wires which were secured radially to the wall of the tank. Four accuracy of ¡0?05% full-scale and 0?5 microstrain resolution.
more layers of sand were deposited, bringing the sand bed to its full An ethernet interface allowed flexible positioning of the sensor
depth of 0?54 m, and producing a pile embedment length of displays on a laptop computer in numeric and graphic modes
0?36 m. This installation scenario represents a ‘wished in place’ using Windows-based ‘Smart-Strain’ software.
closed-ended pile. The test sand had minimum and maximum dry
density values of 1388 and 1662 kg/m3, respectively, which equate 6.4 Testing programme
to maximum and minimum void ratio values of 0?92 and 0?60 A programme of 15 repeating lateral load tests was performed
respectively. This preparation technique produced sand beds with a on the ‘wished in place’ model pile installation in dry sand

62
International Journal of Physical Modelling in Geotechnics Development of a rig to study
Volume 14 Issue 3 model pile behaviour under
repeating lateral loads
Arshad and O’Kelly

One-way
LVDT in vertical direction

LVDTs in horizontal direction


Balanced
Pilehead
Sinusoidal

two-way
50 mm Right load cell
Right spring
Load

Lower right chain segment


Unbalanced 90 mm
two-way Sand bed surface level

Time

One-way Balanced
two-way Figure 5. Arrangement of load cells and LVDTs at the pilehead

Unbalanced
indicates two-way lateral loading, 30 N in one direction, 60 N
two-way
Haversine

in the other (i.e. unbalanced), at a frequency of 0?4 Hz.


Load

The ultimate static lateral load-carrying capacity of the model pile


set-up at the same embedment depth of 0?36 m in the sand beds
was estimated at 140 N, which corresponds to a point on the
Time load–rotation curve where, apparently, plastic deformation of the
surrounding sand starts. At this point, the pile had rotated 1?5˚
Balanced Unbalanced One-way
two-way two-way from its initial vertical alignment and its lateral deflection at the
sand bed surface level was approximately 7 mm; that is, 0?13
times the pile outer diameter D. In previous studies of rigid model
Square

piles performed at 1g, the ultimate load-carrying capacity was


Load

usually estimated for lateral pile deflections of 0?1–0?2D (Cuéllar


et al., 2012; Peng et al., 2011; El Sawwaf, 2006) occurring at the
sand bed surface. In the present study, the load amplitudes of 40
and 60 N investigated correspond to the fatigue and serviceability
Time
limit states, respectively, in relation to the pile’s ultimate static
lateral load-carrying capacity of 140 N (DNV, 2011).
Partial one-way loading

sinusoidal 6.5 Experimental results


Figures 7 to 9 present the experimental results in terms of the
haversine
pile’s rotation from its initial vertical alignment against number
square
of load cycles. The same pile, embedment depth and sand bed
Load

preparation method were used for all of the tests. The


experimental data are discussed below in terms of the effects
Time of loading direction, amplitude and frequency. Using similitude
relationships in the form of non-dimensional parameters (e.g.
see Table 1), such data can be used to interpret the prototype
Figure 4. Loading scenarios and generated wave shapes behaviour. As the model pile was closed-ended and experienced
no vertical loading other than self-weight, its settlement was
negligible, with a maximum of 0?2 mm vertical movement
recorded after 6000 cycles of lateral loading.
beds, prepared as described in Section 6.2. In these tests, the
effects of different loading scenarios (balanced and unbalanced The structural behaviour of a rigid (mono) pile under lateral
two-way and one-way loading), load amplitudes and frequen- loading is defined by its rotation as a rigid body. However, since
cies were investigated, each test involving 6000 lateral loading the pile’s stiffness is not infinite, and provided a high-resolution
cycles applied to the pilehead. The different tests are listed in data-acquisition system such as the System-7000 is used, some
Table 4 and identified as follows: loading direction (1w, one- bending strains at microstrain level can be measured by strain
way; 2w, two-way)/load amplitude (N)/frequency (Hz). For gauges bonded to the pile shaft. These data are usually employed
example, 1w/40/0?1 indicates one-way lateral loading having an for evaluations/interpretations related to the pile’s bending
amplitude and frequency of 40 N and 0?1 Hz; 2w/30–60/0?4 moment profile, the lateral soil reaction and deformations

63
International Journal of Physical Modelling in Geotechnics Development of a rig to study
Volume 14 Issue 3 model pile behaviour under
repeating lateral loads
Arshad and O’Kelly

Test ID Loading scenario Left hanger load: N Right hanger load: N Frequency: Hz

1w/40/0?10 One-way 0 40 0?10


1w/40/0?25 One-way 0 40 0?25
1w/40/0?40 One-way 0 40 0?40
2w/40–40/0?10 Balanced two-way 40 40 0?10
2w/40–40/0?25 Balanced two-way 40 40 0?25
2w/40–40/0?40 Balanced two-way 40 40 0?40
1w/60/0?10 One-way 0 60 0?10
1w/60/0?25 One-way 0 60 0?25
1w/60/0?40 One-way 0 60 0?40
2w/60–60/0?10 Balanced two-way 60 60 0?10
2w/60–60/0?25 Balanced two-way 60 60 0?25
2w/60–60/0?40 Balanced two-way 60 60 0?40
2w/30–60/0?10 Unbalanced two-way 30 60 0?10
2w/30–60/0?25 Unbalanced two-way 30 60 0?25
2w/30–60/0?40 Unbalanced two-way 30 60 0?40

Table 4. Lateral load testing programme on model pile

occurring in the soil strata over the pile embedment length. Figures 8 and 9, it can be interpreted that the sand surrounding
However, to follow the scope of this paper, data obtained from the the pile tended to reach the elastic shakedown state more rapidly
8 strain gauges mounted along the shaft of the model pile over its for the lower LA of 40 N, as compared with 60 N.
embedment depth are not included here and will be used in a
companion paper. 6.5.3 Effect of loading frequency
In relative terms, the effect of loading frequency was more
6.5.1 Effect of loading direction pronounced for the lower LA of 40 N, as compared with 60 N,
Figure 7 considers the same load amplitude of 60 N (service- particularly for one-way loading. For example, between the
ability limit state) and indicates that the measured angle of 500th and 6000th load cycles under one-way loading, the pile
rotation of the pile under unbalanced two-way loading was rotation for 0?4 Hz was approximately 75% greater compared
significantly greater compared with balanced two-way loading. with 0?1 Hz (Figure 8). At the higher LA of 60 N, the
Both of these scenarios were more onerous that one-way corresponding figure was approximately 10%. Under balanced
loading. Similar experimental observations have been reported and unbalanced two-way loading, the respective figures were
from 1g testing of model piles by LeBlanc (2009) and LeBlanc approximately 20% and less than 10% (Figure 9). The increase in
et al. (2010). For these three loading scenarios, approximately pile rotation was not proportionall to the increase in LF; for
80% of the accumulated rotation measured at 6000 load cycles example, under one-way loading, the LF increments from 0?1 to
(i.e. end of the tests) had occurred by the 500th load cycle 0?25 Hz and 0?1 to 0?4 Hz produced increases in pile rotation of
(Figure 7). At 500 load cycles, unbalanced two-way loading 40% and 60% respectively (Figure 8). Under balanced two-way
produced 180% and 75% greater rotation compared with one-way loading, the increase in pile rotation was considerably lower, with
and balanced two-way loading respectively. The corresponding corresponding values of 8% and 22% (Figure 9). For unbalanced
percentages for the 6000th load cycle were 120% and 70%. two-way loading, the LF range investigated was found to have a
negligible effect on the pile rotation response (Figure 7).
6.5.2 Effect of load amplitude
Figures 8 and 9 indicate that the pile rotation was significantly In the existing literature, the effect of LF on accumulated pile
dependent on the load amplitude (LA), particularly for one-way rotation (deflection) is not clear. Our experimental findings are in
loading. For example, with LA increased by 50% (i.e. from 40 to broad agreement with Peng et al. (2011), who reported that up to a
60 N), the pile rotation produced at the 500th load cycle certain limit of LF and for a given LA, pile deflection (lateral
increased by 600% and 100% for one-way and balanced two-way response) increased with increasing LF. Other researchers
loading respectively. At 6000 load cycles, the corresponding (Giannakos et al., 2012; Kagawa, 1986) have reported that at
percentages were 400% and 80%. Based on the calculated higher LF, the surrounding soil becomes stiffer and hence lower
gradients of the data plots between 2000 and 6000 load cycles in deformations are expected under long-term repeating lateral loads.

64
International Journal of Physical Modelling in Geotechnics Development of a rig to study
Volume 14 Issue 3 model pile behaviour under
repeating lateral loads
Arshad and O’Kelly

1 2 3 4 5 6 7 3.0

2.5

2.0

Rotation: degree
8 9 1w/60/0.10
1w/60/0.25
13 1.5 1w/60/0.40
2w/60-60/0.10
14 1.0 2w/60-60/0.25
2w/60-60/0.40
15
16 0.5 2w/30-60/0.10
2w/30-60/0.25
17 2w/30-60/0.40
10 0.0
18 0 2000 4000 6000 8000
11
19 Number of load cycles
12

Figure 7. Effect of loading direction on pile rotation for the same


(a)
load amplitude of 60 N
1

set-ups employing pneumatic or hydraulic actuators, the new


loading system is able to produce similar performance at lower
cost and also provides more control over the waveform shapes.

2
A programme of lateral load tests, each involving 6000 load cycles,
was performed on ‘wished in place’ rigid model piles installed in
dense sand beds. These tests were performed at load amplitudes
(LAs) corresponding to the fatigue and serviceability limit states
10 9 8 7 6 5 4 3 (40 and 60 N, respectively, in the present investigation) to
(b) demonstrate some of the main capabilities of the new loading
system. The following trends were observed from the experimental
Figure 6. Repeating lateral load testing of model pile using the data: (a) two-way loading produced significantly greater rotation
newly developed apparatus: (a) pile in dense sand bed undergoing of the pile from its initial vertical alignment compared with one-
two-way lateral loading; (b) instrumentation at the pilehead. Key way loading; (b) for two-way loading, unbalanced loads produced
(a): 1, sliding node; 2, slot for sliding node; 3, articulated arm; 4, greater pile rotation than balanced loads; (c) the higher LA of 60 N
upper right chain segment; 5, drive chain tensioning device; 6, produced significantly greater pile rotation, particularly for one-
drive motor; 7, upper right cog; 8, laptop computer; 9, System- way loading; (d) the effect of loading frequency was not as
7000 data acquisition system; 10, lower right cog; 11, right load
hanger; 12, right loading weights; 13, upper left chain segment; 1.8
14, vertical LVDT; 15, horizontal LVDT; 16, pilehead; 17, steel tank; 1.6
18, sand bed surface; 19, braced reaction frame. Key (b): 1, vertical 1.4
LVDT; 2, pilehead; 3, sand bed surface; 4, lower right chain
1.2
Rotation: degree

segment; 5, right spring; 6, right load cell; 7, left load cell; 8,


1.0 1w/40/0.10
horizontal LVDTs; 9, left spring; 10, lower left chain segment.
0.8 1w/40/0.25

0.6 1w/40/0.40
1w/60/0.10
7. Conclusion 0.4
1w/60/0.25
A new loading system has been developed to apply many 0.2
1w/60/0.40
thousands of repeating lateral load cycles to a 1g model, with full 0
control provided over the loading direction (i.e. one-way or two- 0 2000 4000 6000 8000
way), amplitude, frequency and waveform shape (e.g. sinusoidal, Number of load cycles
square or haversine). The new loading system is easy to operate
and can create realistic repeating cycles of lateral loadsing. Figure 8. Pile rotation for one-way loading at different load
Hence it is particularly suited to investigations of monopiles amplitudes and frequencies
for offshore structure foundations. Compared with equivalent

65
International Journal of Physical Modelling in Geotechnics Development of a rig to study
Volume 14 Issue 3 model pile behaviour under
repeating lateral loads
Arshad and O’Kelly

1.8
of soil-structure interaction of offshore wind turbines. Soil
Dynamics and Earthquake Engineering 31(5–6): 805–816.
1.6
Bhattacharya S, Lombardi D and Muir Wood D (2011) Similitude
1.4
relationships for physical modelling of monopile-supported
Rotation: degree

1.2
offshore wind turbines. International Journal of Physical
1.0 2w/40-40/0.10 Modelling in Geotechnics 11(2): 58–68.
0.8 2w/40-40/0.25 Bhattacharya S, Cox J, Lombardi D and Muir Wood D (2012)
0.6 2w/40-40/0.40
Dynamics of offshore wind turbines supported on two
2w/60-60/0.10
0.4 foundations. Proceedings of the Institution of Civil
2w/60-60/0.25
0.2 Engineers – Geotechnical Engineering 166(2): 159–169.
2w/60-60/0.40
0 Bhattacharya S, Nikitas N, Garnsey J, Alexander NA, Cox J,
0 2000 4000 6000 8000 Lombardi D, Muir Wood D and Nash DFT (2013) Observed
Number of load cycles dynamic soil–structure interaction in scale testing of
offshore wind turbine foundations. Soil Dynamics and
Figure 9. Pile rotation for balanced two-way loading at different Earthquake Engineering 54(2013): 47–60.
load amplitudes and frequencies Byrne BW and Houlsby GT (2003) Foundations for offshore
wind turbines. Philosophical Transactions of the Royal
Society of London 361(1813): 2909–2930.
significant as loading direction or LA. From the series of tests Chandrasekaran S, Boominathan A and Dodagoudar G (2010)
performed, it was found that the characteristics of the repeating Experimental investigations on the behaviour of pile
lateral loads were more pronounced for the model pile’s fatigue groups in clay under lateral cyclic loading. Geotechnical and
limit state than its serviceability limit state. A more comprehensive Geological Engineering 28(5): 603–617.
programme of repeating lateral load tests is required to further Cuéllar P, Georgi S, Baeßler M and Rucker W (2012) On the
understand the physics behind the problem. quasi-static granular convective flow and sand
densification around pile foundations under cyclic lateral
loading. Granular Matter 14(1): 11–25.
Acknowledgement
Davie JR and Sutherland HB (1978) Modeling of clay uplift
The first author gratefully acknowledges a postgraduate
resistance. Journal of the Geotechnical Engineering Division,
research scholarship award from Trinity College Dublin.
ASCE 104(6): 755–760.
DNV (Det Norske Veritas) (2011) DNV-OS-J101: Design of
REFERENCES Offshore Wind Turbine Structures. DNV, Oslo, Norway.
Achmus M, Kuo Y-S and Abdel-Rahman K (2009) Behavior of Dong P, Newson TA, Davies MCR and Davies PA (2001) Scaling
monopile foundations under cyclic lateral load. Computers laws for centrifuge modelling of soil transport by turbulent
and Geotechnics 36(5): 725–735. fluid flows. International Journal of Physical Modelling in
Adhikari S and Bhattacharya S (2011) Vibrations of wind- Geotechnics 1(1): 41–45.
turbines considering soil-structure interaction. Wind and Dührkop J, Grabe J, Bienen B, White DJ and Randolph FM (2010)
Structures 14(2): 85–112. Centrifuge experiments on laterally loaded piles with wings.
Adhikari S and Bhattacharya S (2012) Dynamic analysis of wind In Proceedings of the 7th International Conference on
turbine towers on flexible foundations. Shock and Vibration Physical Modelling in Geotechnics, Zurich, Switzerland
19(1): 37–56. (Springman S, Laue J and Seward L (eds)). CRC Press/
API (American Petroleum Institute) (2010) Recommended Balkema, Rotterdam, the Netherlands, vol. 2, pp. 919–924.
Practice for Planning, Designing and Constructing Fixed El Nagger MH and Wei JQ (1999) Response of tapered piles
Offshore Platforms — Working Stress Design, API RP 2A- subjected to lateral loading. Canadian Geotechnical Journal
WSD (R2010), 22nd edn. API Publishing Services, 36(1): 52–71.
Washington, DC, USA. El Sawwaf M (2006) Lateral resistance of single pile located near
Arshad M and O’Kelly BC (2013) Offshore wind-turbine geosynthetic reinforced slope. Geotechnical and
structures: a review. Proceedings of the Institution of Civil Geoenvironmental Engineering 132(10): 1336–1345.
Engineers – Energy 166(4): 139–152. Gavin KG and O’Kelly BC (2007) Effect of friction fatigue on pile
Basack S (2005) Development of an apparatus for imparting capacity in dense sand. Geotechnical and Geoenvironmental
lateral cyclic load on model pile foundation. In Proceedings Engineering 133(1): 63–71.
of the International Conference on Mechanical Engineering, Giannakos S, Gerolymos N and Gazetas G (2012) Cyclic lateral
Dhaka, Bangladash, paper AM10. response of piles in dry sand: finite element modeling and
Bhattacharya S and Adhikari S (2011) Experimental validation validation. Computers and Geotechnics 44(2012): 116–131.

66
International Journal of Physical Modelling in Geotechnics Development of a rig to study
Volume 14 Issue 3 model pile behaviour under
repeating lateral loads
Arshad and O’Kelly

Igoe D, Gavin K and O’Kelly B (2013) An investigation into the Peng J, Clarke B and Rouainia M (2011) Increasing the resistance
use of push-in pile foundations by the offshore wind sector. of piles subject to cyclic lateral loading. Geotechnical and
International Journal of Environmental Studies 70(5): 777– Geoenvironmental Engineering 137(10): 977–982.
791. Qin H (2010) Response of Pile Foundations Due to Lateral Force
Kagawa T (1986) Cyclic and loading-rate effects on pile and Soil Movements. PhD thesis, Griffith University,
responses. In Proceedings of the 3rd International Conference Nathan, Australia.
on Numerical Methods in Offshore Piling, Nantes, France. Ramakrishna VGST and Rao SN (1999) Critical cyclic load levels for
Editions Technip, Paris, France, pp. 417–432. laterally loaded piles in soft clays. In Proceedings of the
Klinkvort RT and Hededal O (2013) Lateral response of International Conference on Offshore and Nearshore
monopile supporting an offshore wind turbine. Proceedings Geotechnical Engineering, Panvil, Mumbai, India (Sing SK and
of the Institution of Civil Engineers – Geotechnical Lacasse S (eds)). Balkema, Rotterdam, the Netherlands, pp.
Engineering 166(2): 147–158. 301–307.
Kumar ND and Rao SN (2012) Lateral load: deflection response Rao SN, Ramakrishna VGST and Rao MB (1998) Influence of rigidity
of an embedded caisson in marine clay. Marine on laterally loaded pile groups in marine clay. Geotechnical
Georesources and Geotechnology 30(1): 1–31. and Geoenvironmental Engineering 124(6): 542–549.
Lai S (1989) Simlitude for shaking table test on soil-structure- Rao SN, Ramakrishna VGST and Raju GB (1996) Behavior of pile-
fluid model in 1-g gravitational field. Soils and Foundations supported dolphins in marine clay under lateral loading.
29(1): 105–118. Journal of Geotechnical Engineering, ASCE 122(8): 607–612.
LeBlanc C (2009) Design of Offshore Wind Turbine Support Raongjant W and Meng J (2011) Experimental investigation on
Structures — Selected Topics in the Field of Geotechnical seismic behavior of single piles in sandy soil. Earthquake
Engineering. PhD thesis, Aalborg University, Aalborg, Engineering and Engineering Vibration 10(3): 417–422.
Denmark. Reese LC, Cox WR and Koop FD (1974) Analysis of laterally loaded
LeBlanc C, Houlsby GT and Byrne BW (2010) Response of stiff piles in sand. Proceedings of the 6th Annual Offshore
piles in sand to long-term cyclic lateral loading. Technology Conference, Houston, TX, USA, pp. 473–484.
Géotechnique 60(2): 79–90. Rosquoet F, Thorel L, Garnier J and Canepa Y (2007) Lateral
Lee J, Kim M and Kyung D (2010) Estimation of lateral load cyclic loading of sand-installed piles. Soils and Foundations
capacity of rigid short piles in sands using CPT results. 47(5): 821–832.
Geotechnical and Geoenvironmental Engineering 136(1): 48–56. Sedran G, Stolle DF and Horvath RG (2001) An investigation of
Lee J, Kyung D, Hong J and Kim D (2011) Experimental scaling and dimensional analysis of axially loaded piles.
investigation of laterally loaded piles in sand under Canadian Geotechnical Journal 38(3): 530–541.
multilayered conditions. Soils and Foundations 50(5): 915–927. Tomlinson MJ (2001) Foundation Design and Construction,
Lin SS and Liao JC (1999) Permanent strains of piles in sand due 7th edn. Prentice Hall, Upper Saddle River, NJ, USA.
to cyclic lateral loads. Geotechnical and Geoenvironmental Verdure L, Garnier J and Levacher D (2003) Lateral cyclic
Engineering 125(9): 798–802. loading of single piles in sand. International Journal of
Lombardi D, Bhattacharya S and Muir Wood D (2013) Dynamic Physical Modelling in Geotechnics 3(3): 17–28.
soil–structure interaction of monopile supported wind Zhu B, Byrne BW and Houlsby GT (2013) Long-term lateral cyclic
turbines in cohesive soil. Soil Dynamics and Earthquake response of suction caisson foundations in sand. Geotechnical
Engineering 49(2013): 165–180. and Geoenvironmental Engineering 139(1): 73–83.
Long JH and Vanneste G (1994) Effects of cyclic lateral loads on
piles in sand. Journal of the Geotechnical Engineering WHAT DO YOU THINK?

Division, ASCE 120(1): 225–244. To discuss this paper, please email up to 500 words to the
Mahmoud M and Burley E (1994) Lateral load capacity of single editor at journals@ice.org.uk. Your contribution will be
piles in sand. Proceedings of the Institution of Civil forwarded to the author(s) for a reply and, if considered
Engineers – Geotechnical Engineering 107(3): 155–162. appropriate by the editorial panel, will be published as
Malhotra S (2011) Design and construction considerations for discussion in a future issue of the journal.
offshore wind turbine foundations in North America. In International Journal of Physical Modelling in Geotechnics
Wind Turbines (Al-Bahadly I (ed.)). InTech, Rijeka, relies entirely on contributions sent in by civil engineering
Croatia, pp. 231–264. professionals, academics and students. Papers should be
Muir Wood D, Crewe AJ and Taylor CA (2002) Shaking table 2000–5000 words long (briefing papers should be 1000–
testing of geotectnical models. International Journal of 2000 words long), with adequate illustrations and refer-
Physical Modelling in Geotechnics 2(1): 1–13. ences. You can submit your paper online via www.icevir-
Peng J, Clarke B and Rouainia M (2006) A device to cyclic lateral tuallibrary.com/content/journals, where you will also find
loaded model piles. Geotechnical Testing Journal 29(4): 1–7. detailed author guidelines.

67
Marine Georesources & Geotechnology

ISSN: 1064-119X (Print) 1521-0618 (Online) Journal homepage: http://www.tandfonline.com/loi/umgt20

Analysis and Design of Monopile Foundations for


Offshore Wind-Turbine Structures

Muhammad Arshad & Brendan C. O'Kelly

To cite this article: Muhammad Arshad & Brendan C. O'Kelly (2015): Analysis and Design
of Monopile Foundations for Offshore Wind-Turbine Structures, Marine Georesources &
Geotechnology, DOI: 10.1080/1064119X.2015.1033070

To link to this article: http://dx.doi.org/10.1080/1064119X.2015.1033070

Accepted author version posted online: 28


Jul 2015.
Published online: 28 Jul 2015.

Submit your article to this journal

Article views: 162

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=umgt20

Download by: [University of Engineering & Technology Lahore] Date: 03 February 2016, At: 01:19
Marine Georesources & Geotechnology, 0: 1–23
Copyright # 2015 Taylor & Francis Group, LLC
ISSN: 1064-119X print/1521-0618 online
DOI: 10.1080/1064119X.2015.1033070

Analysis and Design of Monopile Foundations for Offshore


Wind-Turbine Structures
MUHAMMAD ARSHAD1,2 and BRENDAN C. O’KELLY1
1
Department of Civil, Structural and Environmental Engineering, Trinity College Dublin, Dublin, Ireland
2
Department of Geological Engineering, University of Engineering & Technology, Lahore, Pakistan
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Received 10 December 2012, Accepted 19 March 2015

Offshore wind turbines (OWTs) are generally supported by large-diameter monopiles, with the combination of axial forces, lateral
forces, bending moments, and torsional moments generated by the OWT structure and various environmental factors resisted by
earth pressures mobilized in the soil foundation. The lateral loading on the monopile foundation is essentially cyclic in nature and
typically of low amplitude. This state-of-the-art review paper presents details on the geometric design, nominal size, and structural
and environmental loading for existing and planned OWT structures supported by monopile foundations. Pertinent ocean-environ-
ment loading conditions, including methods of calculation using site-specific data, are described along with wave particle kinemat-
ics, focusing on correlations between the loading frequency and natural vibration frequency of the OWT structure. Existing
methods for modeling soil under cyclic loading are reviewed, focusing in particular on strain accumulation models that consider
pile–soil interaction under cyclic lateral loading. Inherent limitations=shortcomings of these models for the analysis and design
of existing and planned OWT monopile foundations are discussed. A design example of an OWT support structure having a mono-
pile foundation system is presented. Target areas for further research by the wind-energy sector, which would facilitate the devel-
opment of improved analyses=design methods for offshore monopiles, are identified.
Keywords: foundation, lateral load, monopile, ocean environment, soil, strain accumulation

Introduction design, the proportion and importance of the different load-


ing types essentially depend upon the kind of foundation
There has been a rapid growth in the use of offshore (and system being considered. For gravity-base foundations
onshore) wind farms for the production of clean and renew- (Figure 1a), potential failure modes may be in bearing
able energy. The economic development of wind farms capacity or excess settlement; hence, the vertical load is gen-
depends on efficient solutions being available for a number erally the major design consideration (Malhotra 2011). For
of technical issues, one aspect being the foundations. Off- monopile foundations (Figure 1b), lateral deflection
shore wind-turbine (OWT) structures may be found on grav- (rotation) of the monopile controls the serviceability limit
ity base, suction caisson, monopile, tripod or braced frame state of the whole structure; hence, lateral loads are more
(jacket) foundations or, more recently, using floating plat- critical when compared with the vertical loads.
forms tethered to the seabed (Figure 1). The foundation In general, offshore foundations must be designed to
choice largely depends on the water depth, seabed character- resist large numbers of aerodynamic and hydrodynamic load
istics, loading characteristics, available construction technol- cycles of varying direction, amplitude, and frequency
ogies, and importantly, economic costs (Malhotra 2011). (Figure 2) at the proposed site over the project’s lifetime of
Offshore foundations are subjected to a combination of typically 25 years or more (Şahin 2004). In most parts
loads, namely the axial (self-weight) forces of the of the world, the most popular foundation choice in terms
structure=machinery (V), repeating horizontal=lateral loads of ease of installation, economy, and logistics is the mono-
(H), and bending (M) and torsional moments. Apart from pile, with an estimated 75% of all installed OWTs using this
the self-weight, these loads are generated by environmental solution (Blanco 2009; Fischer 2011; Malhotra 2011). For
conditions (ocean waves, currents, tidal action, and aerody- OWT monopile foundation systems, the lateral loads (and
namic load cycles) and=or operation of the installed machin- resulting moments) are generally large in proportion to the
ery (Figure 2). The lateral loading is essentially cyclic in axial loads. Hence, the foundation response under repeating
nature and typically of low amplitude. For geotechnical lateral loading is a major consideration, with the foundation
design dominated by considerations of the dynamic and fati-
Address correspondence to Muhammad Arshad, Department gue responses under working loads, rather than the ultimate
of Geological Engineering, University of Engineering & Tech- load-carrying capacity. The overall feasibility of a wind-farm
nology, GT Road Lahore, Pakistan. E-mail: arshadm@tcd.ie project is invariably determined by upfront capital costs,
2 M. Arshad & B. C. O’Kelly
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Fig. 2. Environmental impacts on offshore wind turbine.


Fig. 1. Support structure options showing their range of appli-
cable water depths.
generation capacity, which is an indirect measure of the
with the material, fabrication, and installation costs for applied loading. Other significant influencing factors include
OWT foundations, typically accounting for 25% of the the soil condition and severity of environmental loading.
overall project cost (Blanco 2009). This significant invest- Existing offshore monopiles are typically 5–7.5 m in diam-
ment demands reliable foundation designs to achieve the eter (Achmus, Kuo, and Abdel-Rahman 2009), with slender-
overall goals of pollution-free wind-generated power at ness [embedded length (L) to outer diameter (D)] ratio
economical rates. values of <10, and are; therefore, considered to behave as
The aims of this review article are to present: rigid structures for which rotation is prominent over bending
(Tomlinson 2001; Peng, Clarke, and Rouainia 2011). For a
. Details on the geometric design and structural and typical OWT having a rated power of 5 MW and installed
environmental loading for existing and planned OWTs in the North Sea, a hub height (see Figure 3) of 95 m above
supported by monopile foundations. the mean sea level and rotor diameter of 125 m lead to the
. Describe ocean-environment loading conditions, including approximate quasi-static loading scenario summarized in
methods of calculation using site-specific data. Table 1 (Lesny and Wiemann 2005; IEC [International
. Discuss wave particle kinematics, focusing on correlations Electrotechnical Commission] 2009) and which acts at the
between the loading frequency and natural vibration fre- seabed level. This example demonstrates that torsional
quency of OWT structures. moments are negligibly small, whereas lateral loading causes
. Critically review existing semi-empirical methods for mod- extremely high bending moments and principally controls
eling soil under cyclic loading, focusing on cyclic design the foundation design.
for fatigue life. The inherent limitations=shortcomings of The monopile foundation transfers the horizontal forces
these models for the analyses=design of existing and and bending moments generated by wind and wave loading
planned OWT monopile foundations are identified along (via horizontal earth pressure acting along the pile) into the
with further research needs. surrounding ground through cantilever action. Factors
. Present a typical example of the geometric design for an affecting the cyclic response include the monopile size and
OWT-monopile foundation system. embedment length, soil properties, soil–pile relative stiffness,
Other aspects related to offshore monopile structures loading characteristics, and pile installation method (Malho-
(e.g., construction materials, manufacturing advancements, tra 2011). Depending on subsoil conditions, the monopile is
corrosion assessment, and control) are beyond the scope of typically installed into the seabed by drilling and grouting,
this article. driving using large impact or vibratory hammers, or a com-
bination of drilling and driving (Malhotra 2011).
In current practice, OWT monopile foundations are
General Aspects of OWT-Monopile Foundation usually designed using general geotechnical standards in
System combination with more specific guidelines and semi-
empirical formulas that have been largely developed by the
A monopile is a single large-diameter steel tube that pene- offshore oil=gas industry. The latter are usually based on
trates into the seabed. The required monopile length and limited field data obtained for piles having significantly
diameter are primarily dependent on the OWT’s power smaller diameters (DIN [Deutsches Institut fur Normung]
Analysis and Design of Offshore Monopile Foundations 3

2.0 and 1.4 MN, respectively: these scenarios correspond to


the serviceability and fatigue limit states respectively (GL
2005). Further, the strategy of predictions and in-situ
measurement of the behavior of every OWT monopile foun-
dation in a wind farm project is technically and economically
laborious and; hence, not practical. The application of fast
countermeasures to mitigate against unexpected conditions
developing into potentially alarming situations also cannot
be guaranteed (Richwien, Lesny, and Wiemann 2002).
Many constitutive models have been proposed to estimate
the soil response under large numbers of load cycles, although
very few of these specifically consider simulation of the soil–
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

pile interaction, instead focusing on the soil’s strength and


stiffness responses. The most popular method of analysis for
predicting the pile deflection under lateral loading, and the
method recommended in offshore design codes including
API (2010) and DNV (2011), are based on the Winkler model;
commonly referred to as the p-y method. Although it has a
long history of use in design practice, the original database
on which this method is based largely consists of a single set
of pile tests, involving limited numbers of lateral load cycles,
performed at a medium-dense sand site. In contrast, many
millions of low-amplitude cycles of lateral loading are applied
over the operational life of OWT structures.
As a general rule, under field loading, rotation of the
monopile of up to 0.5 from its vertical alignment (Achmus,
Kuo, and Abdel-Rahman 2009; LeBlanc, Houlsby, and
Byrne 2010; DNV 2011) or lateral pile deflections (based
on practical experience) of up to 120 mm occurring at seabed
level (Malhotra 2011) are considered as limiting values for
the proper operation of the wind turbine.

Geometric Details of OWTs Supported by Monopile


Foundations
The main components of the wind-turbine system are its
foundation, support structure, transition piece, tower, rotor
Fig. 3. Main components of offshore wind turbine system. blades, and nacelle (Figure 3). Together, the support struc-
ture and foundation maintain the turbine in its proper oper-
2005; GL [Germanischer Lloyd] 2005; API [American ational position. The transition piece provides a means of
Petroleum Institute] 2010; DNV [Det Norske Veritas] correcting any misalignment of the monopile that may arise
2011). Hence, extrapolation of these formulations for during its installation. In some instances, the monopile can
present and future design of OWT monopile foundations extend from the seabed level to above the water surface level,
requires careful consideration of the applied loading and connecting with the transition piece or tower (Arshad and
the inherent limitations underlying such approaches. O’Kelly 2013). The nacelle (hub) contains the key electro-
Another significant shortcoming is that current design mechanical components of the wind turbine, including the
standards and guidelines on the serviceability of piles under gearbox and generator (Malhotra 2011). Rotor diameters
cyclic lateral loading are limited. For instance, over its ser- of existing OWTs typically range 90–120 m, producing
vice life, a typical 2 MW OWT structure is usually subjected power-generation capacities of 3–6 MW respectively (Tong
to 102 and 107 cycles of lateral loading with magnitudes of 2010). Gravitational loading typically ranges 2–8 MN. For
instance, Byrne and Houlsby (2003) reported vertical load
Table 1. Loading at seabed level for 5 MW wind turbine of 6 MN for an anticipated 3.5 MW OWT in the UK region,
supported by monopile foundation (Lesny and Wiemann 2005) with lateral loading from aerodynamic and hydrodynamic
Component Magnitude factors accounting for up to 67% of the vertical loading.
These loads will vary with the size of the installation, the
Vertical load 35 MN detailed design, and local environmental conditions. OWTs
Horizontal load 16 MN having rotor diameters of 250 m and a power-generation
Bending moment 562 MN  m capacity of 20 MW are currently in the research and develop-
Torsional moment 4 MN  m ment phase (EWEA [European Wind Energy Association]
4 M. Arshad & B. C. O’Kelly
 a
2011), creating even more onerous design loading scenarios Z
for (monopile) foundations. U z ¼ U 10 ð1Þ
10
OWT monopile foundations are typically manufactured
from steel tubular sections having outer diameters of up to ln Z=Z0
7.5 m, wall thicknesses of up to 150 mm, and embedment U z ¼ U 10 ð2Þ
ln 10=Z0
(penetration) depths of between 15 and 30 m (Achmus,
Kuo, and Abdel-Rahman 2009). Monopiles are generally where the values of exponent a and the roughness parameter,
used in shallow water depths (i.e., typically <30 m), gener- Z0, are dependent on the site location, with typical values of
ally becoming too flexible for water depths between 30 0.11 and 0.0001, respectively, for open-sea environments.
and 40 m, in which case, monopiles fitted with guy wires or The wind velocity fluctuates with time for a given height,
tripod solutions (see Figure 1) are considered as economical and also with change in height, above the sea level or ground
alternatives. For greater depths, time-consuming installation surface (Figure 4). However, a mean value (U z) can be estab-
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

and the effects of soil degradation (‘potholing’), which lished for a particular site location provided sufficient wind
occurs at seabed level around the pile, make monopile foun- speed data are available. The instantaneous wind velocity
dation solutions prohibitive (Irvine et al. 2003). Other foun- can be considered as the superposition of the turbulent
dation options, as illustrated in Figure 1, are then considered component (wz) on the mean value. Hence the total aerody-
as viable options. The serviceability limit state is largely namic drag force (FDWind) acting on an offshore structure
determined by the lateral deflection (rotation) of the mono- can be determined by (Jang and Shinn 1999; API 2010):
pile under many millions of load cycles; e.g., 107 lateral load  2  
cycles of 1.4 MN magnitude (corresponding to the fatigue FDWind ¼ 0:5qair Cdair Aw U z þ qair Cdair Aw U z wz
loading for design) are expected to occur over the service life þ 0:5qair Cdair Aw ðjwz jwz Þ ð3Þ
of OWT structures (GL 2005).

Ocean-Environment Loading on Monopile Foundation


System
In the ocean-environment, the primary sources of variable=
cyclic loading on OWTs are aerodynamic, hydrodynamic,
and (depending on the geographic setting) ice loading.
Seismic loads (considered as a special type of dynamic load-
ing) and ship impact may threaten the serviceability of the
structure or even ultimately lead to its collapse. However,
the focus of this paper is on aerodynamic and hydrodynamic
loading, since these are considered more important in assess-
ments of the OWT structure’s fatigue life.

Aerodynamic Loading
Wind conditions are important in defining, not only the
loads imposed on a wind turbine’s structural components,
but also in predicting the amount of future energy produced
as a function of wind velocity. A realistic assessment of wind
direction through statistical analysis of recorded wind data
must be based on a realistic representation of wind speed
(preferably occurring at hub height), speed frequency distri-
bution, wind shear (i.e., rate of change in wind speed with
height), turbulence intensity (i.e., standard deviation of wind
speeds sampled over a 10 min period as a function of the
mean speed), wind direction distribution, and also extreme
wind gusts with return periods of up to 100 years (DNV
2011). The mean value of the 10 min period wind speed data
measured at a reference elevation of 10 m above mean sea
level (usually determined at hub height for OWTs) is referred
to as the wind speed U 10 , from which the mean wind speed
U z for some other height, Z, above mean sea level can be
approximated using either the power law or logarithmic Fig. 4. Fluctuation of wind speed about the mean value. (a) In
law given by: time domain. (b) In three-dimensional space.
Analysis and Design of Offshore Monopile Foundations 5

where qair is the density of air, Aw is the exposed area of the


offshore structure, Cdair is an aerodynamic drag coefficient
that is dependent on the structure’s geometry and size, jwzj
is the turbulent component (with the absolute value taken
in order to make the direction of the drag force coincident
with the wind velocity).
The first term on the right-hand side of Eq. (3) represents
the mean drag force, with the second and third terms denot-
ing the drag force components associated with wind velocity
fluctuations.
Wind loads on offshore structures are seldom greater than
the hydrodynamic loads and often play a relatively minor
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

role, compared with the hydrodynamic loads. However wind


loading has a longer lever arm when considering moments
generated about the structure’s foundation. Hence, it is best
not to neglect the effects of wind loading, particularly in the
calculation of overturning (bending) moments acting on the
foundation. Wind loads on offshore structures are seldom
greater than the hydrodynamic loads. However, in the calcu-
lation of the overturning (bending) moments acting on the
foundation, wind loading has a longer lever arm compared
with the wave load interaction with the OWT structure. For
instance, Byrne and Houlsby (2003) reported that in a typical
North Sea environment, the rotor thrust reaction due to the
wind loads contributes 25% of the total horizontal load
but generates 75% of the total overturning moments.
Variations in mean wind speed are relatively small com-
pared with variations in the wave period. Fluctuations about
the mean wind speed impose aerodynamic forces, although
when considering the dynamic behavior of offshore struc-
tures, these forces are also generally insignificant compared
with the hydrodynamic forces (Journée and Massie 2001).
Hence wind speed is generally considered as steady (i.e., with- Fig. 5. Definitions of harmonic wave propagation and para-
out fluctuation), resulting in constant force(s) and; hence, meters. (a) In time domain. (b) In three-dimensional space.
constant moment(s). Dynamic analysis of offshore structures
is necessary in cases where the wind field contains energy at useful in many applications since it allows one to predict
frequencies near the offshore structure’s natural frequencies complex irregular behavior in terms of the theory of regular
of vibration (API 2010), although for monopile foundations, waves; e.g., see St. Denis and Pierson (1953).
the difference between these frequencies is usually high [see
later section on Correlation between Loading Frequency Wave Particle Kinematics
and Natural Frequency of OWT Structures]. Some statistical Wave Theory
manipulations can be applied to the wind data (e.g., refer to
Jang and Shinn 1999; Haritos 2007) for the inclusion of Many regular wave theories have been developed to capture
dynamic forces due to wind velocity turbulence. the water particle kinematics associated with ocean waves.
Each theory pertains to a certain scenario with different
degrees of complexity. Hence, they have different ranking
Hydrodynamic Loading
among the offshore engineering community (Chakrabarti
For OWT monopile foundations, the dominant hydrody- 2005). Airy wave theory is the simplest of the routinely-used
namic loads are generated by ocean waves and currents, with regular wave theories. Its key features are introduced in this
minor contributions arising from other sources such as sea paper in the context of describing its role in modeling the
level variations due to tides or swell. The main sources of character of irregular sea states. Referring to Figure 5, with
wave generation are wind and astronomical forces. Earth- the wave moving in the positive horizontal direction (x), the
quakes, submarine landslides, and shipping or other nearby form of the water surface (wave profile) can be expressed as
floating structures may also be important sources of wave a function of x and time (t) by:
generation. Irregular waves are usually observed on the
ocean surface, although these can be treated as a superpo- f ðx; tÞ ¼ fa cos ðkx  xtÞ ð4Þ
sition of many regular harmonic waves, each having its and with the wave moving in the opposite direction by:
own amplitude (na), wavelength (k), frequency (f), and direc-
tion of propagation (see Figure 5). This concept can be very f ðx; tÞ ¼ fa cos ðkx þ xtÞ ð5Þ
6 M. Arshad & B. C. O’Kelly

where k is the wave number, x is the circular wave In the time domain, the total horizontal load per unit
frequency, and fa is the maximum amplitude of the water length [F(t)] exerted on a fixed object (e.g., OWT monopile
particle measured from the mean water surface level. foundation) as a result of wave motion and currents can
The resulting water particle velocity components u(t) and be considered as the linear addition of the inertial [Finertia(t)]
v(t) in the respective x and vertical directions can be and drag [Fdrag(t)] forces given by Morison, Johnson, and
expressed by: Schaff (1950):
coshðkhw þ kzÞ F ðtÞ ¼ Finertia ðtÞ þ Fdrag ðtÞ
uðtÞ ¼ fa x cos ðkx  xtÞ ð6Þ
sinhðkhw Þ
sinhðkhw þ kzÞ p
vðtÞ ¼ fa x sin ðkx  xtÞ ð7Þ Finertia ðtÞ ¼ q CM D2  u_ ðtÞ ð10Þ
sinhðkhw Þ 4 w
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

where hw is the mean depth of the water body and z is the


elevation of the water particle above the mean surface level 1
Fdrag ðtÞ ¼ qw CD D  uðtÞ  juðtÞj ð11Þ
in the vertical direction. 2
For deep water conditions, these velocity equations
where qw is the density of sea water, D is the cylinder (i.e.,
reduce to:
monopile) diameter over the distance between the seabed
and ocean surface, and u(t) ¼ uacos(xt), ju(t)j and u_ ðtÞ are
uðtÞ ¼ fa xekz cos ðkx  xtÞ ð8Þ
the horizontal water particle velocity, its absolute value
vðtÞ ¼ fa xekz sinðkx  xtÞ ð9Þ and its time derivative (acceleration), respectively, with ua
equal to the maximum along wave water-particle velocity
From Eqs. (9) and (10), a simple interpretation is that for [discussed earlier in section on Wave theory], CM and CD
deep water conditions, the elliptical orbits of water particles are the inertial and hydrodynamic drag force coefficients,
associated with Airy wave theory reduce to circular orbits. respectively, whose values are dependent on Reynold’s num-
Different methods are available to determine the horizontal ber (Re) and Keulegan–Carpenter’s number (KC):
velocity profile with depth, including Wheeler profile stretch-
ing, extrapolation, and constant extension methods. Since ua D
the wave loads are dependent on the water particle velocity, Re ¼ ð12Þ
W
horizontal wave loads for shallow waters may be scrupu-
lously undervalued if the calculations are based on deep ua T
water equations. When considering maximum velocity KC ¼ ð13Þ
D
values, the time function of Eq. (6) can be ignored without
any significant loss in accuracy. The time derivatives of where W is the kinematic viscosity and T is the oscillating
Eqs. (6–9) give the respective accelerations, which are key wave period.
parameters for the estimation of the inertial forces acting A correct evaluation of the total hydrodynamic load act-
on offshore structures located in the flow field. ing on an offshore structure must consider the combined
current flow and wave particle velocities. Wave particle velo-
cities and accelerations are sinusoidal in time, and when seen
Morison’s Equation
as functions of time, are out of phase by 90 from each other.
Morison, Johnson, and Schaff (1950) formulated an equa- Adopting the absolute value of the velocity term (i.e., ju(t)j)
tion to predict the wave forces acting on a vertical pile in Eq. (11) synchronizes the drag force and flow direction.
exposed to horizontal sinusoidal oscillatory flow. In their Irregular variations of the total hydrodynamic load are pro-
equation, the linear inertial force and adapted quadratic duced against time for an idealized record, as demonstrated
drag force (from real flows and constant currents) are super- by the experimental data shown in Figure 6 for a monopile
imposed to obtain the resultant force acting on the projected at reduced laboratory scale. For inertia dominated wave
area. For larger offshore structures (e.g., gravity foundations loads and (or) small current velocity, the loads arising from
and also large diameter monopiles), the wave field is signifi- the ocean current can be ignored (DNV 2011).
cantly influenced and a diffraction regime emerges. The The dominancy of the drag force or inertial force has been
potential flow theory is more suitable for the calculation of assessed in terms of the Keulegan–Carpenter number. Typi-
wave loads on such structures (Batchelor 1967). Morison’s cal recommendations give the inertial force dominating for
formula is strictly limited for use with slender structural ele- KC < 10, comparable inertial and drag forces for
ments; i.e., characterized by D=k < 0.2: where D is the diam- 10 < KC < 20, with the drag force dominating for KC > 20.
eter of the structural element between the seabed level and In general, the CM and CD coefficients are dependent on
the transition piece (see Figure 3) and k is the impinging both the Reynold’s and Keulegan–Carpenter’s numbers
wavelength of the ocean wave. However a significant num- (Haritos 2007). For OWT monopile foundations, widely
ber of existing offshore structure designs have employed used design codes such as API (2010) and DNV (2011)
Morison’s equation even though the criterion of D=k < 0.2 specify=recommend appropriate CD and CM values, depend-
may not have always been fully satisfied (Haritos 2007). ing on the pile surface roughness and (or) the Re number.
Analysis and Design of Offshore Monopile Foundations 7
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Fig. 7. Structural model of flexible wind-turbine system.

equivalent steel pipe for the support structure. In this case,


the first natural frequency (fnat, in Hz) of the whole structure
can be approximated by (Tong 2010):

2 3:04 EI
fnat ¼ ð14Þ
4ðpÞ ð0:227 lLi þ Mt Þ ðLi Þ3
2

where l is the structure mass to length ratio (kg=m); Li is the


structure height=(strut length) and EI is its bending stiffness
(Nm2).
Offshore wind-turbine OWT structures are excited by
Fig. 6. Graphical representations of variations in (a) velocity wind and waves, with the effective wind load determined
and acceleration, and (b) horizontal force against time for an by a complex interaction between the structural dynamics
idealized record related to a monopile at reduced laboratory of the turbine and the wind field, which contains turbulent
scale (Morison, Johnson, and Schaff 1950). gusts caused by eddies in the flow. The turbulent wind field
originates from atmospheric turbulence, with nearby
turbines also contributing to the disturbance of the flow.
Correlation between Loading Frequency and Natural
Energy-rich wind turbulence occurs at below 0.1 Hz
Frequency of OWT Structures
(LeBlanc 2009). The height of the ocean waves is usually
It is essential to consider the fundamental natural frequency defined in terms of ‘significant wave height’ and determined
of an OWT structure for a proper description and evaluation as the mean value of the highest one-third of the waves in a
of its dynamic behavior. For dynamic systems, resonance given wave record. Ocean waves that induce fatigue loading
occurs if an excitation frequency gets close to the structure’s with high frequency on offshore structures usually have sig-
natural frequency. For OWTs, this invariably leads to the nificant wave heights ranging 1.0–1.5 m and also have a
development of higher stresses and, more significantly, to a zero-crossing period (Tz) of 4–5 s (De Vries 2007).
higher range of stresses in the support structure, which is Modern OWTs are installed with either pitch-regulated
an unfavorable situation with respect to fatigue life. Hence blades or variable rotational-speed systems to enable optimi-
it is important to ensure that excitation frequencies having zation of their power production over a wide range of wind
high energy levels do not coincide with the support struc- speeds. The rotational speed of the main rotor shaft is typi-
ture’s natural frequency. As a first approximation, the sup- cally in the range 10–20 rpm. Hence the first excitation fre-
port structure’s natural frequency can be determined by quency ‘1P’ (i.e., corresponding to one full revolution)
considering a simplified geometry for the whole structure occurs in the range 0.17–0.33 Hz and, in general, should only
(Figure 7). The turbine mass (Mt) is concentrated at the top be lightly excited. Large excitations in the 1P frequency
(free end), similar to a cantilevered strut, which represents an range generally arise from excessive turbine mass or
8 M. Arshad & B. C. O’Kelly

comprehensive literature supports inherent anisotropic


behavior (or the general cross-anisotropy present in many
sedimentary deposits) for the soil’s strength=deformation
characteristics and also seepage properties (Yasin and
Tatsuoka 2000; Wang and Lade 2001; O’Kelly 2005, 2006).
Due to the initial and stress-induced anisotropy of most
deposits (Naughton and O’Kelly 2004), soil deformations
can occur whenever the magnitudes of the three principal
stresses change and (or) the principal stress axes reorientate
on account of the cyclic lateral loading. An analogy may be
drawn between the behavior of the soil in the immediate
vicinity of an OWT monopile foundation and that of a
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

pavement under repeated wheel loading of varying intensity.


For both scenarios, changes in both the magnitudes and the
directions of the principal stresses occur in a single cycle.

Fig. 8. Excitation frequency ranges for offshore wind turbines Simulation of In-situ Stress Conditions in the Geotechnical
having rated power-generation capacities ranging 2.0–3.6 MW Laboratory
(LeBlanc 2009). (a) Deflected shape of monopile. (b) Soil press-
The values of pertinent parameters used to describe the soil
ure pt exerted due to pile deflection yt for a specific depth xt. (c)
Winkler model approach and change in shape of p–y curves
response under cyclic loading can be determined in the geo-
with depth. technical laboratory using cyclic triaxial tests (Das 2008),
although the system of cyclic axial loading and lateral con-
finement pressure acting on the test-specimen is axisym-
aerodynamic imbalances. For a three-bladed turbine, the metric. An advancement on the cyclic triaxial apparatus is
blade passing-frequency of typically 0.5–1.0 Hz is denoted the hollow cylinder apparatus (HCA) (O’Kelly and
by the ‘3P’ frequency, which is heavily excited, mainly on Naughton 2005a) which allows independent control of the
account of the impulse-like excitation arising from the blades magnitudes of the three principal stresses and also the orien-
passing by the tower. Site-specific spectral densities for wind tation of the major–minor principal stress axis. The HCA is
and waves can be derived from measured site data, met-ocean ideal for simulating cyclic multi-directional loading con-
databases or using numerical models (LeBlanc, Houlsby, and ditions on cross-anisotropic test specimens. Generalized
Byrne 2010). Figure 8 illustrates the 1P and 3P excitation stress-path testing can be performed in which the stress his-
ranges, along with realistic normalized power-spectra tory and in-service loading conditions at specific locations in
representing aerodynamic and hydrodynamic excitations. the soil foundation can be simulated under stress- (O’Kelly
The regions before the 1P frequency range and after the and Naughton 2009) or strain-controlled conditions
3P frequency range are referred to as the ‘Soft–Soft’ and (Naughton and O’Kelly 2005; Das 2008). Special prep-
‘Stiff–Stiff’ zones, respectively. The structure will be too flex- aration techniques (O’Kelly and Naughton 2005b) are
ible if its natural frequency falls within the ‘Soft–Soft’ zone required to prepare=reconstitute the test specimen (which
and too rigid (heavy and expensive) if its natural frequency is often disturbed during the sampling procedure) in a phy-
falls within the ‘Stiff–Stiff’ zone; both of these scenarios sically identical condition to that of the in-situ deposit. In
making it unsuitable for the design. Wind and wave-turbu- many practical situations, laboratory testing may become
lence excitation frequencies usually fall within the ‘Soft–Soft’ too laborious, expensive, and (or) time consuming. In-situ
zone: another important reason for avoiding this frequency testing techniques, including the Cone Penetration Test
region (LeBlanc, Houlsby, and Byrne 2010). method (Igoe et al. 2013) are used for offshore site investiga-
tions and afford another approach in the determination of
the pertinent design parameter values.
Modeling Soil Behavior Under Cyclic Loading
Real Soil Behavior Modeling Soil Under Cyclic Loading
Soil deposits can be classified in many ways; e.g., by forma- The strain accumulation occurring in soil under repeated
tion process, grain size (fine or coarse), plasticity index, age loading is dependent on the material properties, stress
(recent or aged deposits), mineralogical content, etc. Apart path=level, and number of load cycles (Niemunis,
from at very small strain levels of <103 strain (Atkinson Wichtmann, and Triantafyllidis 2005; Karg 2007). Many
and Sallfors 1991; O’Kelly and Naughton 2008), the stress– models with different complexity and acceptability have
strain relationship for soil is generally highly nonlinear been developed for the prediction of the strain accumulation
(inelastic) (Budhu 2011), with the strength and stiffness occurring in a soil element under cyclic loading. These
properties strongly dependent on stress history, drainage include models that are broadly based on: the number of
conditions (drained or undrained) and the stress path load cycles (Barksdale 1972; Sweere 1990; Hornych, Corte,
followed during loading. For undisturbed deposits, and Paute 1993); the stress level (Paute, Hornych, and
Analysis and Design of Offshore Monopile Foundations 9

Benaben 1996); the number of load cycles and the stress level in-service over many decades. However, caution is necessary
(Pappin 1979; Lentz and Baladi 1981; Li and Selig 1996; in applying this methodology to OWT monopile foundation
Lekarp and Dawson 1998; Chai and Miura 2002) or the design, since the approach may often be applied outside of
number of load cycles, stress level and material properties its verified range and several important design issues may
(Niemunis, Wichtmann, and Triantafyllidis 2005; Karg not be properly considered. The API (2010) and DNV
2007). A major limitation to their application in offshore (2011) standards rely on methods (models) built upon
foundation design calculations is that none of these models empirical data obtained for long flexible piles, for which
explicitly consider the (mono)pile–soil interaction under lat- bending (deflection) is significant. In contrast, existing and
eral loading. In reality, the pile deflection (rotation) response planned OWT monopile foundations invariably have slen-
under lateral loading arises from the soil behavior, which is derness ratios of <10 (typical range of 5–6), indicating rigid
dependent on the loading conditions. A few models have pile behavior (Achmus, Kuo, and Abdel-Rahman 2009;
been tailored for this particular scenario; these are described Peng, Clarke, and Rouainia 2011). Under these circum-
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

in the next section, in the context of the design of OWT stances, the pile rotation is generally more prominent over
monopile foundations. A discussion on numerical modeling, bending, with the rotation occurring about a point (‘axis
considering dynamic constitutive soil models, torsional load- of rotation’) located approximately one pile diameter above
ing, and damping related issues, is beyond the scope of this the pile base. Hence the design criteria and analyses appro-
review article. Details on these topics can be found in Basack priate for flexible and rigid piles are considerably different
and Dey (2012), Basack and Sen (2014), Guo (2006, 2013) (Dobry et al. 1982), casting doubt on the application of
and Rani and Prashant (2014). the API (2010) and DNV (2011) methods based on p–y
curves in predicting the in-service behavior of offshore
monopiles. Further, the p–y curves for cyclic loading pre-
Strain Accumulation Models Considering Pile–Soil sented in API (2010) and DNV (2011) were primarily formu-
Interaction lated for the evaluation of the ultimate lateral load-carrying
capacity mobilized under relatively few load cycles. In con-
p–y Model trast, OWT monopile foundations experience many millions
In general terms, a p–y curve is typically obtained by plotting of low-amplitude cycles over their in-service life. Further, the
the soil pressure (p) response against the pile’s lateral
deflection (y) arising from the action of a horizontal load
(H) applied at the pile head (Figure 9a). Figure 9b shows
the soil pressure (pt) distribution generated around the pile
circumference at a particular depth (xt) and the correspond-
ing pile deflection (yt) response. In the literature curves (e.g.,
Matlock 1970; Reese, Cox, and Koop 1975; Ismael 1990; API
2010), p–y curves can be categorized on the basis of soil type
(granular or cohesive), loading type (monotonically increas-
ing or cyclically repeated) and the groundwater table level.
The effects of soil stratification, nonlinearity, and other
soil properties are automatically considered by determining
p–y curves specific to different depth ranges along the length
of the pile (Figure 9c), which is typically modeled using
Winkler’s approach; i.e., the pile member acts as a beam
supported by a series of uncoupled nonlinear elastic springs
that represent the soil reaction. For instance, soil stiffness
generally increases with depth (overburden pressure), which
is reflected by increasing values of the spring stiffness (Epy),
defined as the secant modulus of the p–y curve (Figure 9c).
The pile deflection that develops under given loading con-
ditions and constrains can be predicted by implementing
the related p–y curve in a simple non-dimensional frame-
work, such as Randolph’s method (Randolph 1981), or by
numerical methods using computer software such as
COM624P (1993). However, despite the stiffness Epy being
a soil–structure interaction parameter, API (2010) and
DNV (2011) only consider the soil properties in formulating
the p–y curves, and the influence of the pile properties on the
mobilized p–y curves remains an open question.
Current pile design methodology based on p–y curves, as
described in API (2010) and DNV (2011), has gained
broad recognition on account of the low failure rate for piles Fig. 9. p–y method of analysis for laterally loaded pile.
10 M. Arshad & B. C. O’Kelly

API (2010) and DNV (2011) methods do not provide a determined by the pile length and the relative stiffness Tr
means of calculating the accumulated pile deflection [¼(EI=nhN)0.2] value.
(rotation) that occurs during cyclic loading. Changes in the In Eq. (16), all of the variables can be taken as known
foundation stiffness as a result of long-term cyclic loading, except for coefficient nhN, which is calculated as:
which typically densifies (but in some circumstances may
loosen) the surrounding soil (LeBlanc, Houlsby, and Byrne nhN ¼ nh1 N t ð17Þ
2010), are also poorly accounted for in current design meth-
where nh1 is the coefficient of soil reaction for the first cycle
odologies based on p–y curves.
of loading, its value dependent on the relative density of the
soil and the location of the groundwater table (Terzaghi
Model of Little and Briaud (1988) 1955; Reese, Cox, and Koop 1974), with the exponent t (hav-
The Little and Briaud (1988) model is based on experimental ing a recommended range of 0.2–0.4 (Broms 1964; Davisson
1970)) determined empirically by:
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

results obtained from six cyclic lateral load tests performed


on long pre-stressed concrete and steel pipes (D ¼ 0.6–1.06 m)
installed in a 22 m thick stratum of loose to medium dense sand, t ¼ 0:17 FL Fi FD ð18Þ
overlying stiff clay. Twenty load cycles were performed for each where FL, Fi, and FD are empirically determined factors that
test pile. The following power function was proposed for esti- take into account the influence of the cyclic load ratio, pile
mating the accumulated lateral strain (en) developed at ground installation method, and soil density, respectively. Note
surface level for N number of load cycles: nh1 ¼ nh; i.e., its value is the same as that for the monotonic
en ¼ e 1 N m ð15Þ (static) loading condition.
The Long and Vanneste (1994) model provides a rela-
where e1 is the strain response at the end of the first load cycle tively straightforward procedure for predicting the effect of
and exponent m is a parameter that accounts for the influence cyclic lateral loading, but specifically restricted to scenarios
of the soil properties, pile material, installation method, and involving the soil reaction modulus increasing linearly with
loading characteristics, with experimentally derived m values depth. Soil stratification, nonlinearity, and other fundamen-
ranging 0.04–0.09 for the ‘flexible’ test piles investigated (Little tal parameters (e.g., unit weight and strength) that affect the
and Briaud 1988). In this context, the lateral strain is calculated lateral load response are not explicitly taken into account by
as the pile’s lateral deflection occurring at the seabed level div- this model. Further, most of the thirty-four pile tests
ided by its outer diameter; a dimensionless quantity. considered in the development of Eq. (16) involved the
Numerical studies by Kuo, Achmus, and Abdel-Rahman application of less than fifty lateral load cycles.
(2012) indicated an m value of 0.07 for ‘flexible’ piles (agree-
ing with the experimental range reported by Little and Briaud Model of Lin and Liao (1999)
(1988)) and 0.135 for ‘rigid’ piles. The validity of Eq. (15) has Based on the results of twenty-six full-scale piles tested under
been verified using cyclic lateral load test data obtained for cyclic lateral loading (N < 50), Lin and Liao (1999) proposed
model piles installed in medium-dense quartz sand (Peralta a logarithmic trend to capture the accumulated strain en
and Achmus 2010). It has also been shown that the value of developed at the ground surface level for N load cycles:
exponent m can be evaluated from consolidated–drained cyc-
lic triaxial test data (Peralta and Achmus 2010). Nevertheless, en ¼ e1 ½1 þ t lnðN Þ ð19Þ
Eq. (15) is empirical and formulated on limited test data where t is a degradation parameter determined by:
(N < 20) for a loose to medium dense sand deposit.
Model of Long and Vanneste (1994) L
t ¼ 0:032 FL Fi FD ð20Þ
The Long and Vanneste (1994) model is based on a closed- Tt
form solution for a pile in an elastic foundation soil, with where Tt ¼ ðEI=nh Þ0:2 , L is the embedded length of the pile
the soil reaction modulus increasing linearly with depth. and nh is the coefficient of soil reaction for the monotonic
By applying this approach and using input parameters (static) loading condition.
derived from reported field data for cyclic lateral load tests The other contributing factors in Eq. (20) are similar to
performed on thirty-four piles of different lengths, dia- those in Eq. (18), with insignificant differences in values
meters, materials, and installation techniques, Long and from those reported for the Long and Vanneste (1994)
Vanneste (1994) proposed that the accumulated lateral strain model. However, considering the various influencing factors,
occurring at ground surface level for N load cycles can be caution is urged regarding empirical parameters used in
estimated by: describing the logarithmic change in the pile deflection=
AH BM rotation.
en ¼ þ ð16Þ In their study, Lin and Liao (1999) considered two load-
EI0:4 ðnhN Þ0:6 EI0:6 ðnhN Þ0:4
ing scenarios: one-way loading, in which the applied hori-
where EI is the flexural rigidity of the pile; nhN is the zontal load increases from zero to its maximum value
coefficient of soil reaction for the Nth cycle of loading; H (Hmax) before reducing back to zero, thereby completing
and M are the horizontal load and moment, respectively, one cycle of loading; two-way loading, in which the horizon-
applied at the ground surface level; A and B are scalars tal load is applied in the opposite direction for half of each
Analysis and Design of Offshore Monopile Foundations 11

model explicitly considers the effects of soil density, mean


effective confining pressure, cyclic stress amplitude, simul-
taneous oscillation of stresses from different directions,
and the number of load cycles:

e_acc ¼ fampl f_N fe fp fY fp ð21Þ


where e_acc is the derivative of strain accumulation with
respect to the number of load cycles N, fampl is the strain
amplitude function, f_N is the load-cycle number function,
fe is the soil density function, fp is the mean effective confin-
ing pressure function, fY is the average stress ratio (i.e.,
mobilized deviatoric stress to average mean confining press-
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

ure applied during a load cycle) function and fp is the polar-


ization (load cyclic shape) function. Further details on the
determination of these functions and commonly adopted
values are reported by Wichtmann, Niemunis, and Trianta-
Fig. 10. Different modes of load variation against time.
fyllidis (2009), Wichtmann et al. (2010).
Wichtmann et al. (2010) applied the high cycle accumulation
cycle, reducing its value below zero to a minimum of Hmin model for typical North Sea fine sand deposits and found that
(see Figure 10). greater permanent (plastic) lateral deflection occurred for:
According to the Long and Vanneste (1994) and Lin and
. Decreasing relative density, with the average value of the
Liao (1999) models, one-way loading produces greater accu-
mulated strain compared with two-way loading. Similar applied moment (Mave) and its amplitude (Mampl) held
overall behavior has been reported from studies on model constant;
. Increasing Mampl, with the relative density and Mave held
monopiles (Peng, Clarke, and Rouainia 2011), with balanced
two-way loading (Hmax ¼ Hmin) producing about 80% of constant;
. Increasing Mave, with the relative density and Mampl held
the lateral deflection recorded under one-way loading of
the same magnitude. Unbalanced two-way loading constant.
(Hmax 6¼ Hmin) was reported to produce even lower These outcomes are qualitatively conceivable and broadly
amounts of lateral deflection. However some recent physical consistent with the behavior observed in laboratory studies
model studies suggest that unbalanced two-way loading is on model monopiles by LeBlanc, Houlsby, and Byrne
more damaging to the lateral stability of the monopile (2010) and Peng, Clarke, and Rouainia (2011), although
(Zhu, Byrne, and Houlsby 2013; Arshad and O’Kelly further physical model studies and in-situ testing are
2014). An increase in soil stiffness was also observed with required to confirm these findings. Potentially significant
increasing number of load cycles. Similar formulations to shortcomings of the high cycle accumulation model include:
that given by Eq. (19) have been proposed by Verdure, the multiplicative approach used in Eq. (21) means that
Garnier, and Levacher (2003) and Li, Haigh, and Bolton potentially small errors in the values of the individual func-
(2010) based on centrifuge experiments performed on minia- tions may produce an objectionable error in the final calcu-
ture piles installed in sandy deposits, with the degradation lation value; the formulation includes no measure of loading
parameter t estimated by curve-fitting analysis of the experi- frequency which is an important factor in connection with
mental data. The change in soil stiffness was found highly the natural vibration frequency of OWT structures [see
dependent on the cyclic load amplitude. Up to certain levels Section on Correlation between Loading Frequency and
of cyclic load amplitude and numbers of load cycles, the stiff- Natural Frequency of OWT Structures] and the strain
ness was found to increase rapidly, but no further change was accumulation rate.
observed for higher load amplitudes, even during the first few
cycles (Verdure, Garnier, and Levacher 2003). Model of Achmus, Kuo, and Abdel-Rahman (2009)
The Achmus, Kuo, and Abdel-Rahman (2009) model is
High Cycle Accumulation Model (2005, 2010) based on the concept of the soil stiffness degradation
The high cycle accumulation model is based on the experi- approach suggested by different researchers (e.g., Little
mental data obtained from a large collection of experimental and Briaud 1988; Long and Vanneste 1994), with emphasis
work, comprising index tests and drained cyclic triaxial, res- on the degradation of the soil secant stiffness (Es), in keeping
onant column and simple shear tests performed on sand spe- with formulations proposed by Huurman (1996) and Werk-
cimens having a wide range of initial relative densities meister (2003). According to this model, an increase in the
(Niemunis, Wichtmann, and Triantafyllidis 2005). Although plastic strain can be interpreted as a decrease in the Es value.
this model was initially developed without consideration of In cyclic triaxial tests (for which elastic strains are usually
pile–soil interaction, its potential application in modeling a negligible) and referring to Figure 11, the plastic axial strains
pile–soil system under low-amplitude stress cycles has been produced at the end of the first and Nth cycles (epN ¼ 1 and
demonstrated by finite-element calculations using ABAQUS epN ¼ N) can be determined from the rate of degradation of
software (Wichtmann et al. 2010). The formulation of this the secant modulus measured for the first and Nth cycles
12 M. Arshad & B. C. O’Kelly
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Fig. 11. Degradation of soil secant modulus (Es) under cyclic loading.

(EsN ¼ 1 and EsN ¼ N respectively) by: and accounts for the loading and unloading phases emerging
epN¼1 due to the cyclic horizontal loading. Achmus, Kuo, and
EsN¼N
¼ ð22Þ Abdel-Rahman (2009) found good agreement between
epN¼N EsN¼1 numerical simulations performed using cyclic triaxial test
data (Timmerman and Wu 1969) and experimental data
In the Achmus, Kuo, and Abdel-Rahman (2009) study, for laterally loaded model monopiles installed in sand.
the characteristics of the permanent strains derived from
Model of LeBlanc (2009)
drained cyclic triaxial test data for medium and dense sand
were implemented in a 3D finite-element model of a laterally The LeBlanc (2009) model is based on lateral load tests,
loaded pile. The Achmus, Kuo, and Abdel-Rahman (2009) involving between 8  103 and 6  104 load cycles, performed
model suggests that the accumulated strain en developed on rigid model piles having a slenderness ratio of 4.5 and
for N load cycles can be estimated using the semi-empirical which were installed in unsaturated, very loose and medium
approach proposed by Huurman (1996) for the calculation dense Leighton Buzzard sand beds. A non-dimensional
of the accumulated plastic strains in cyclic triaxial tests, as: framework was developed to identify realistic pile dimen-
sions and loading ranges for the laboratory study. Some
e1 particular loading characteristics were found to cause signifi-
en ¼ ð23Þ
b1 ðxc Þb2 cant increases in the accumulated pile rotation to occur, with
ðN Þ
the soil stiffness increasing with the number of load cycles.
where b1 and b2 are model parameters and xc is the charac- This contrasts starkly with the Long and Vanneste (1994),
teristic cyclic stress ratio (ranging 0–1 (Achmus, Kuo, and API (2010) and Achmus, Kuo, and Abdel-Rahman (2009)
Abdel-Rahman 2009)), defined as: models for which the static load–displacement curves are
ðcyclic stress ratio at loadingÞ degraded to account for cyclic loading.
LeBlanc (2009) proposed that the accumulated rotation hN
 ðcylic stress ratio at unloadingÞ (and hence the lateral displacement) of a ‘rigid’ pile developed
xc ¼
1  ðcylic stress ratio at unloadingÞ for N load cycles can be estimated by (see Figure 12):

The cyclic stress ratio at constant confining pressure is hN ¼ h0 þ DhðNÞ ð24Þ


defined as the ratio of the major principal stress under cyclic where h0 is the pile rotation achieved when the applied
loading to the major principal stress at failure under static moment reaches its maximum value during the first load cycle
loading. Since the horizontal loading varies cyclically for and Dh(N) is given by:
OWT monopile foundations, the confinement pressure act-
ing on soil elements surrounding the pile shaft also changes.
DhðN Þ ¼ hs Tb ðfb ; Rd ÞTc ðfc Þ N 0:31 ð25Þ
Hence the cyclic stress ratio values for cyclic triaxial tests (in
which the axial load is varied cyclically) and OWT monopile where hs is the pile rotation produced by a static load having
foundations (where the axial load remains constant and the the same magnitude as the maximum cyclic load (Figure 12b),
horizontal load varies cyclically) are different. To overcome Tb and Tc are dimensionless functions, Rd is the relative
this, the ‘characteristic cyclic stress ratio’ is used in Eq. (23) density of the sand, with fb (a measure of the cyclic load
Analysis and Design of Offshore Monopile Foundations 13
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Fig. 13. Functions relating (a) Tb and (b) Tc to the cyclic load
Fig. 12. Variation in pile rotation h with fluctuation in moment
characteristics in terms of fb and fc respectively.
M. (a) Cyclic test (b) Static test.

amplitude normalized with respect to the static moment Further, this model cannot account for the multidirectional
capacity, MR) and fc (quantifies characteristics of the cyclic loading typical of offshore wind-farm environments.
load) given by: Considering the accumulated strain en produced at the
ground surface level for a rigid pile is proportional to the
Mmax accumulated rotation, then the predictions by this model
fb ¼ ð26Þ (with due consideration for the cyclic character of the load-
MR
ing) are contrary to those for flexible piles according to the
Mmin Long and Vanneste (1994) and Lin and Liao (1999) model.
fc ¼ ð27Þ For loose and medium dense sand, LeBlanc (2009) also
Mmax
reported an increase in soil stiffness due to cyclic loading
where Mmin and Mmax are the minimum and maximum of the rigid pile, although further research for higher densi-
moments developed in a load cycle. fication levels, other loading frequencies and degrees of satu-
It follows that 0 <fb < 1, with fc equal to unity for a static ration are necessary to confirm this preliminary experimental
load test, zero for one-way loading and 1 for balanced two- finding.
way loading. Values of Tb and Tc, determined from experi-
mental data of Dh(N)=hs against N, were plotted against fb Model of Bienen et al. (2012)
and fc, as shown in Figure 13. Note that with Tc ¼ 0 for The Bienen et al. (2012) model is based on experimental data
fc ¼  1.0, Eq. (25) proposes that the rotation occurring from miniature monopiles installed in dry medium-dense
after the first load cycle is equal to zero for balanced two- sand and tested at 1g and 200g, using a centrifuge apparatus
way cyclic loading, which is fairly unrealistic in practice. for the latter. The test piles were representing a prototype
14 M. Arshad & B. C. O’Kelly

pile 2.4 m in diameter (D), with embedment lengths (L) of 9.6 Design Procedure for OWT Foundation System
and 30 m; i.e., 4D (‘rigid’) and 12.5D (flexible) piles, respect-
ively. Using a sinusoidal waveform, different magnitudes of Design Motive
one-way lateral loading were investigated considering 104 The basic driving motive for the design procedure is to avoid
load cycles and a single model frequency of 0.25 Hz. The the occurrence of resonance in the dynamic behavior of the
accumulated lateral deflection y occurring at the pile head structural system under in-service loading (Jaimes 2010). An
(assumed flush with the surrounding ground surface) for N iterative procedure is usually adopted in design, with the
load cycles can be estimated by: basic steps involved in the design process for an OWT mono-
 as pile foundation system illustrated in Figure 14. The data
H
y ¼ D fN As 100 2 ð28Þ required include environmental, turbine and site stratigra-
D LQc phy (soil profile) data. The environmental data is used to
where H is the horizontal load applied to the pile head, Qc is determine the required work-platform and hub elevations
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

the rate of change in CPT cone-tip resistance with depth, As (see Figure 3) for the proposed OWT, and to select initial=
and as are dimensionless parameters dependent on the soil trial dimensions=geometry for the monopile foundation,
properties and pile dimensions (Dyson and Randolph leading to the determination of the natural frequency of
2001; Dührkop 2009), with reported values of As ¼ 1.4 and the whole structural system. Checks on resonance frequency
as ¼ 0.0072 for monopiles (Bienen et al. 2012), and fN (factor are applied along with predictions of the anticipated
to modify the monotonic response (deflection) into a cyclic rotation, deflection, and settlement responses of the pro-
response) approximated by: posed foundation system produced by the applied loads=
moments that are determined from the environmental and
N 1 wind turbine data. Fatigue and buckling checks are per-
fN ¼ 1 þ BN1 ðlnðBN2 þ 1Þ ð29Þ formed at a more advanced stage of the design process,
N
usually using some computer software package, and hence
where the parameters BN1 and BN2 are determined from a are beyond the scope of this paper. The following sections
plot of [(y=D)n=(y=D)1] against number of load cycles N. present a design example in which the work-platform and
This model, based on the one-way lateral loading and sin- hub elevations along with the required embedment length
gle frequency of 0.25 Hz investigated, was found to provide for a monopile foundation supporting a typical 5 MW
reasonable predictions of the accumulated strain produced OWT (Jonkman et al. 2009) are determined.
under cyclic loading, although the soil stiffness response
was not clear: increasing for up to certain levels of accumu-
lated strain, but then decreasing with further load cycles. Determination of Work Platform and Hub Elevations
Hence these findings should be verified for greater numbers
Referring to Figure 14, the first step in the design process
of load cycles occurring in multiple directions and for other
requires the determination of the turbine platform and hub
frequency ranges, degrees of soil saturation and densification
elevations, usually with reference to the lowest astronomical
levels.
tide (LAT); defined as the lowest tide level to occur under
Model of Klinkvort and Hededal (2013) average meteorological conditions and any combination of
Klinkvort and Hededal (2013) applied the model proposed astronomical conditions. LAT is below the mean sea level
by LeBlanc (2009) to experimental data for ‘rigid’ model (MSL). The high astronomical tide (HAT) is above the
piles (slenderness ratio of 6) installed in saturated and dry MSL, but below the storm surge level.
dense sand that were tested at ng using a centrifuge appar- The work-platform level is located at the top of the
atus. These tests involved a maximum of 500 load cycles transition piece and it determines the location of the flange
and were performed at the same frequency and soil density.
They proposed that the pile’s accumulated lateral strain pro-
duced at the ground surface level after N number of load
cycles (en) can be estimated as:
Tb ðfbb ÞTc ðfcc Þ
en ¼ e1 N ð30Þ
with fbb and fcc defined as:

Pmax
fbb ¼ ð31Þ
Pu
Pmin
fcc ¼ ð32Þ
Pmax
where Pu is the ultimate lateral load-carrying capacity under
monotonic loading applied to the pilehead; Pmax and Pmin
are the maximum and minimum lateral loads, respectively, Fig. 14. Flowchart for the design of offshore wind-turbine
applied to the pilehead during the cyclic loading. support structure and monopile foundation.
Analysis and Design of Offshore Monopile Foundations 15
Table 2. Values of key parameters for typical 5 MW wind Determining the Required Natural Frequency and
turbine Preliminary Tower Dimensions
Parameter Magnitude From the rotational speed interval of the main rotor shaft,
as given in Table 2, the 1P region ranges 0.12–0.20 Hz and
Turbine mass (rotor and nacelle) 350 tonnes the 3P region ranges 0.36–0.60 Hz (Jonkman et al. 2009).
Rotor diameter 126 m For this particular design example, the target value of the
Nominal rotor speed 10.1 rpm natural frequency [Eq. (14)] was set as 0.27 Hz; i.e., within
Rotational speed interval 6.9–12.1 rpm the ‘wanted frequency’ range of 0.20–0.36 Hz. By taking into
of the main rotor shaft
account the hub height, wind-turbine mass and target natu-
Cut-in wind speed 4 m=s
ral frequency, the outer diameter and wall thickness at the
Nominal wind speed 11 m=s
top of the steel tower were estimated as 3.9 m and 40 mm,
Cut-out wind speed 25 m=s
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

respectively, with corresponding values at the tower base


of 6.6 m and 100 mm. These preliminary selected dimensions
connection between the transition piece and the turbine must also satisfy the buckling and fatigue checks; otherwise
tower (see Figure 3). The magnitudes of the applied wind the calculations are revised iteratively. The outer diameter of
loads and the location of the centre of gravity of the nacelle the transition piece (DTP) subsequently depends on the diam-
mass is strongly dependent on the hub height. Representa- eter of the monopile, as:
tive values for key parameters of the typical 5 MW wind tur-  
bine considered in this design example are given in Table 2. DTP ¼ D þ 2 tTP þ tgrout ð36Þ
Environmental data characteristic of an offshore wind-farm where D is the monopile diameter (usually equal to the tower
located in the North Sea are given in Table 3. For this base diameter), tTP is the wall thickness of the transition
location, a tidal range (DZtide ¼ HATLAT) of 1.6 m and piece and tgrout is the thickness of grout connection.
storm surge (DZsurge) of 2.0 m may be adopted (Jonkman
et al. 2009; Fischer 2011).
The required work-platform level (Zplatform) is given by: Calculation of Design Loads
The design loads for the subject monopile foundation
Zplatform ¼ LAT þ DZtide þ DZsurge þ DZair þ n ð33Þ system were estimated using the preliminary tower geometry,
the wind-turbine parameter values given in Table 2 and the
where DZair is the air gap of typically 1.5 m between the
environmental data given in Table 3. For this example, the
work-platform level and the highest wave crest elevation (n).
software package ‘RECAL’ was used to estimate the shear
For an assumed wave-height coefficient value of 0.65 and
force and bending moment acting on the monopile foun-
a maximum wave height, relative to the MSL, having a 50
dation at seabed level, as documented by De Vries and
year return period (Hmax(50 year)) of 20.3 m, the value of
van der Tempel (2007). The torsional moments are usually
n is given in DNV (2011) as:
only a small fraction (<1%) of the bending moment (Lesny
n ¼ 0:65 Hmax ð50 yearÞ  13:2 m ð34Þ and Wiemann 2005); e.g., see typical loading values listed in
Table 1 for a 5 MW wind turbine having a monopile foun-
And hence the value of Zplatform is determined as
dation system. Hence, provided checks on lateral forces
LAT þ 19 m (rounding the latter up to the nearest whole
and bending moments are satisfied, torsional moments are
integer). A check should also be applied (De Vries and van
generally not critical and have not been specifically con-
der Tempel 2007 DNV 2001) such that Hmax(50
sidered in design examples for OWT monopile foundation
year) < 0.78  30 m; the assumed MSL value for these calcula-
systems reported in the literature. RECAL can simulate both
tions. The required hub height (Zhub) of LAT þ 87m above
wind and wave time-series, from which it calculates the
seabed level is determined from:
environmental loads acting on the support structure, as well
Zhub ¼ Zplatform þ 0:5ðrotor diameterÞ þ 5 ð35Þ as the structure’s dynamic behavior. For the design example
under consideration, Table 4 presents load data in terms of
With LAT ¼ MSL  Ztide=2, Zhub ¼ 116.2 m from seabed the lateral forces and bending moments generated by
level. RECAL. The reported load values take into account the
safety factors for hydrodynamic loading and aerodynamic
Table 3. Extreme wind velocity (Vw), current velocity (Uc),
significant wave height (Hs), and maximum wave height (Hmax)
as a function of return period (Treturn) for reference offshore Table 4. Estimated design loads occurring at seabed level for
wind-farm site (Jonkman et al. 2009) design example

Treturn (year) Vw (m=s) Uc (m=s) Hs (m) Hmax (m) Lateral force, H Bending moment, M
Load type (MN) (MN.m)
1 33.9 0.70 7.72 14.35
5 38.0 0.80 9.03 16.80 Aerodynamic 1.42 127
10 39.8 0.84 9.60 17.85 Hydrodynamic 9.27 298
50 43.9 0.94 10.91 20.29 Total 10.69 425
16 M. Arshad & B. C. O’Kelly

thrust on the rotor, in addition to the different critical design example are given in Figure A1 of the Appendix. It
combinations of wind and wave loading, as reported in was assumed that the monopile was made of steel having
several guidelines=standards (API 2010; DNV 2011). yield stress and Young’s modulus values of 248 MPa and
207 GPa, respectively.
The embedment length of the monopile was initially
Monopile Embedment Length and Foundation Stability
assumed equal to nine times its outer diameter. Checks on
The monopile’s embedment length must be sufficient to the pile rotation and (or) its deflection occurring at the
ensure vertical and lateral stability. Studying the interaction seabed level are applied and the monopile embedment length
effects for piles under combined axial and lateral loading is optimized accordingly. The monopile’s rotation (h0) and
would call for a systematic and sophisticated analysis, lateral deflection (v0) occurring at the seabed level under
although the pertinent literature is limited and sometimes the design loads (horizontal force (H) and bending moment
contradictory. For instance, analytical studies by Ramasamy (M)) are calculated by considering the closed-form solutions
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

(1974) and Goryunov (1973) indicate that for a given lateral proposed by Randolph (1981), Broms (1964) and Matlock
load, the pile’s lateral deflection increases for the combi- and Reese (1960). These methods are related to monotonic=
nation with vertical loading but some experimental static loading conditions and hence there is no consideration
(Sorochan and Bykov 1976; Jain, Ranjan, and Ramasamy of the number of load cycles. Randolph’s (1981) method,
1987) and field (Bartolomey 1977; Zhukov and Balov illustrated in Figure 15, employs Eqs. (37) and (38) to calcu-
1978) studies have indicated the contrary. Numerical analy- late v0 and h0, respectively; with parameters Ep (effective
sis by Karthigeyan, Ramakrishna, and Rajagopal (2007) Young’s modulus of the pile), Gc (equivalent shear modulus
indicated that for sandy soils, the presence of vertical loads of the soil at a depth of 0.5Lc), Lc (active pile length) and Rc
increases the pile’s lateral load-carrying capacity by as much (the soil’s equivalent shear modulus profile parameter)
as 40% (depending on the magnitude of the axial loading), defined in Figure 15. In this method, the monopile dimen-
although marginal reductions in the lateral load-carrying sions are incorporated in terms of its active length and
capacity were found to occur for clayey soils. second moment of area, I; the latter an indirect involvement
According to current practice, for monopile design, separ- of its cross-sectional area (i.e., its inner and outer diameter
ate analyses are performed that consider (i) the axial loading dimensions). When the active length exceeds the actual=pro-
only, to determine the bearing capacity and settlement posed monopile embedment length (L), the monopile tends
response, and (ii) the lateral loading only, to determine the
flexural behavior through cantilever action (Karthigeyan,
Ramakrishna, and Rajagopal 2006; Moayed, Mehdipour,
and Judi 2012; Rahim and Stevens 2013). However, com-
pared with the axial loads, the lateral loads are considered
governing, as mentioned in several design guidelines (e.g.,
GL 2005; API 2010; DNV 2011) and documented by many
researchers (Achmus 2010; LeBlanc, Houlsby, and Byrne
2010; Malhotra 2011; Peng, Clarke, and Rouainia 2011;
Bhattacharya et al. 2012; Kuo, Achmus, and Abdel-Rahman
2012; Haiderali, Cilingir, and Madabhushi 2013; Lombardi,
Bhattacharya, and Wood 2013; Zhu, Byrne, and Houlsby
2013; Nicolai and Ibsen 2014; Carswell et al. 2015). For
OWT monopile foundations, the pile must mobilize suf-
ficient soil resistance over its embedded length to transfer
all the different types of applied loads to the surrounding
soil, with adequate safety factors, and prevent toe ‘kick’ (dis-
placement at the pile base) and excessive deflection=rotation
of the pile itself from occurring. The pile size and embed-
ment length necessary to satisfy the lateral load requirements
are generally greater compared with those necessary to
satisfy the axial loading requirements (Kopp 2010). Hence
this design example focuses on the lateral stability of the sub-
ject monopile.
The input data for the monopile design includes the soil
strength and stiffness (e.g., elastic (Es) and shear (Gs) modu-
lii and Poisson’s ratio) profiles against depth, the pile
properties (e.g., cross-sectional dimensions=properties and
material strength=stiffness), and the design loads (De Vries
2007; Jaimes 2010; Tong, 2010). The soil (medium dense
sand considered for the present case) properties used in the Fig. 15. Closed-form solution after Randolph (1981).
Analysis and Design of Offshore Monopile Foundations 17

to translate somewhat in the soil foundation as well as embedded length of the pile. However, for a ‘rigid’ pile, this
deflecting; hence the monopile deflection obtained through decreasing trend may not be applicable. According to this
Eq. (37) must be modified according to Eq. (39). method (Matlock and Reese 1960), beyond a certain value
 0:142  1  2 ! of embedment length, there will be no further reduction in
Ep =GC Lc Lc the monopile’s deflection or rotation. Calculations are not
v0 ¼ 0:27H þ 0:3M ð37Þ
Rc Gc 2 2 included for the present design example, but the pile embed-
ment length should be sufficient to achieve tolerable values
 0:142  2  3 ! of v0 and h0 occurring at the pile toe level. For the calcu-
Ep =GC Lc 0:5 Lc
h0 ¼ 0:3H þ 0:8ðRc Þ M lation of the v0 and h0 values occurring at the seabed level,
Rc Gc 2 2
a parametric study was performed, varying the pile’s embed-
ð38Þ ment length, outer diameter and inner diameter in the ranges
  20–40 m, 5.5–7.0 m, and 5.4–6.5 m, respectively. A summary
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Lc
v0 ðadjusted valueÞ ¼ 0:8v0ðfrom Equation 37Þ: ð39Þ of the results of this study is presented in Figures 17–19 and
L conversed as:
. The monopile deflection=rotation values predicted by the
Broms’s (1964) method of calculation for v0 is illustrated three selected methods for the static design load conditions
in Figure 16. Matlock and Reese’s (1960) method employs grossly differed from one another; e.g., for the applied
Eqs. (40) and (41) to calculate the values of v0 and h0, loads and range of monopile embedment lengths and
respectively, occurring at the seabed level.
HTt 3 MTt 2
v0 ¼ Av0 þ Bv0 ð40Þ
EI EI
HTt 2 MTt
h0 ¼ Ah0 þ Bh0 ð41Þ
EI EI
where Av0, Bv0, Ah0, and Bh0 are sets of non-dimensional
coefficients (scalars) whose values are dependent on the
depth along the embedded length of the pile and Tt ¼ (EI=
nh)0.2, with the value of the coefficient of soil modulus vari-
ation, nh, dependent on the soil relative density and the
location of the groundwater table (Terzaghi 1955; Reese,
Cox, and Koop 1974).
For the seabed (ground surface) level, the values of coeffi-
cients Av0, Bv0, Ah0, and Bh0 are 2.43, 1.62, 1.62, and 1.75,
respectively, with these coefficients decreasing in value with
increasing depth along the pile embedment length, reflecting
the reduction in v0 and h0 as we move down along the

Fig. 16. Charts for calculating the lateral deflection occurring at Fig. 17. Using Randolph’s (1981) method, effects of pile’s
the ground surface level for laterally loaded piles in cohesionless embedment length and diameter on its (a) lateral deflection
soil (after Broms (1964)). occurring at the seabed level and (b) rotation.
18 M. Arshad & B. C. O’Kelly

minor change in v0 was observed for pile embedment


lengths greater than 26 m, irrespective of the pile’s cross-
sectional area (Figure 18).
. For the Matlock and Reese’s (1960) method, v0 and h0
were both found dependent on the inner and outer dia-
meters (cross-sectional area) of the monopile, although,
their significance tended to reduce substantially for pile
inner and outer diameters greater than 5.8 and 6.0 m
respectively (Figure 19).
The monopile’s lateral response for the long-term cyclic
lateral loading condition was assessed for the presented
design example using the models after Little and Briaud
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

(1988), Long and Vanneste (1994), Lin and Liao (1999)


and Klinkvort and Hededal (2013), which have been dis-
cussed previously in this paper. It was assumed that the
monopile had been driven into the medium dense sand
deposit and was subjected to one-way cyclic loading.
Fig. 18. Summary of the results deduced using Broms’s (1964)
Figure 20 presents the accumulated lateral response of the
method.
monopile predicted by the different models considered.
The accumulated lateral response is in fact a multiple of
inner=outer diameters considered in the design example, the response for the end of the first load cycle, realistic values
the predicted values of v0 ranged 70–270, 25–115 and of which cannot be estimated using the models considered,
25–90 mm for the Randolph’s (1981), Broms’s (1964) and for the purpose of these calculations, its value was
and Matlock and Reese (1960) methods respectively. This assumed equal to unity. Hence Figure 20 should be viewed
noticeable difference is perhaps due to the involvement of in the context of a qualitative comparison of the long-term
marked empiricism in the methods used to evaluate the predictions by the different models. From Figure 20, it can
lateral stability of the different monopile set-ups and the be concluded that for the same loading conditions, large dif-
limited available experimental data used in their ferences can occur between the accumulated lateral
formulation=calibration. responses of the monopile predicted by the different models,
. For Randolph’s (1981) method, v0 was found sensitive to perhaps by many folds for large numbers of load cycles. One
the pile’s embedment length and its inner and outer dia- reason for these large differences can be put down to the
meters (incorporated in calculations of the cross-sectional empirical selection of values for the coefficients=non-dimen-
area and second moment of area, I) (Figure 17a), whereas sional parameters incorporated in these models for the pre-
h0 was sensitive to the pile cross-sectional area only diction of the monopile’s lateral response under long-term
(Figure 17b). cyclic loading.
. For Broms’s (1964) method, v0 was found dependent on The monopile’s ultimate axial load-carrying capacity
the embedment length and inner and outer diameters (under static loading) (Qd) can be determined using
(cross-sectional area) of the monopile, although only a

Fig. 19. Summary of the results deduced using the Matlock and Fig. 20. Accumulated lateral strain responses predicted for the
Reese’s (1960) method. monopile at seabed level under long-term cyclic loading.
Analysis and Design of Offshore Monopile Foundations 19

Eq. (43), which is the recommended practice by API (2010), . Pile properties (e.g., diameter and slenderness ratio) and
and has been employed by many researchers; e.g., De Vries the soil–pile interaction;
(2007); Igoe, Gavin, and O’Kelly (2010); Haiderali, Cilingir, . Varying the direction and frequency of the applied lateral
and Madabhushi (2013); Bisoi and Haldar (2014); to name loading;
a few. . Combined (e.g., axial and lateral) cyclic loading.
The scarcity of field data for cyclic lateral loading of large
Qd ¼ Qf þ Qb ¼ fAsur þ qAb ð42Þ
diameter (rigid) piles available in the literature, particularly
where Qf is the shaft-friction capacity, Qb is the end-bearing for high load cycling, makes the validation of current and
capacity, f is the unit skin friction, Asur is the shaft area improved design methods=theories and the calibration of
pertaining to the pile embedment length, q is the unit end- numerical models for offshore monopiles difficult. Many
bearing capacity, and Ab is the base area of the plugged of the proposed models=formulations, with the principal
ones described earlier in the paper, were calibrated against
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

monopile.
For monopiles having a diameter of 4 m diameter or experimental data obtained from model-scale pile tests.
greater, the pile plug resistance is usually not taken into Further instrumented field testing of full-scale monopiles,
account in the calculations (van der Tempel 2006). Further, ideally having comparable size and geometry (slenderness
it has been shown that degradation of the shaft-friction ratio) with that of current and proposed OWT monopile
capacity due to cyclic axial loading leads to accumulating foundations, and subjected to high numbers of lateral load
displacements and potentially a severe reduction in the axial cycles, is warranted and would provide a valuable source
load-carrying capacity (Gavin and O’Kelly 2007), although of information in this regard for the wind-energy sector.
consideration of this degradation of the shaft friction in
the design process is still an open question. Numerous design
Summary and Conclusions
charts available in the literature make it possible to dis-
tinguish between stable and unstable loading levels to ensure Sources, types, and methods of analyses for the determi-
a design solution on the safe side. Further, no reduction in nation of the magnitude of the environmental loading and
axial load-carrying capacity is expected if a certain magni- resulting moments exerted on offshore monopiles, including
tude of the cyclic load amplitude is not exceeded (see Poulos for long-term and extreme conditions, are well documented
1988, and Abdel-Rahman and Achmus 2011, for further in the literature. However the behavior of the pile–soil sys-
details). tem in response to these cyclic (dynamic) loading scenarios
is not entirely clear. Further, the analysis=design of large
diameter (rigid) monopile foundations for current and pro-
Discussion posed OWT structures is well outside the scope of present
experience and analysis=design methods, which are mainly
There is no overall agreement in the literature regarding the based on experimental data obtained for relatively small-
determination of the pile’s rotation=lateral deformation diameter flexible piles subjected to low numbers of load
response to the many millions of low-amplitude lateral load cycles (N < 200). This includes the widely used API (2010)
cycles associated with OWT monopile foundation scenarios. and DNV (2011) standards developed primarily for the off-
Considerable differences of opinion exist on the rate of cyclic shore oil=gas industry.
strain accumulation; e.g., power function (Little and Briaud The behavior of large-diameter monopiles in general, and
1988) and logarithmic-trend relationships (Lin and Liao the long-term low-amplitude cyclic lateral loading response
1999) have been proposed. Most of the reported field and in particular, are not well documented. Existing models for
laboratory pile tests were performed for medium dense sand estimating the accumulated lateral strain (rotation) response
(Little and Briaud 1988; LeBlanc 2009; Bienen et al. 2012). of monopiles are based on very limited field=laboratory data
Compared with the Bienen et al. (2012) and LeBlanc and are currently not capable of explicitly accounting for
(2009) models which are based on N  104 lateral load cycles, site-specific soil properties and environmental loading char-
other widely used models for predicting the pile’s rotation= acteristics. There is also no consensus among researchers
lateral displacement response, including the API (2010), regarding the severity of strain accumulation due to one-
DNV (2011), Little and Briaud (1988) and Long and way and two-way loading scenarios, the effects of varying
Vanneste (1994) approaches, are based on experimental data the load direction and frequency or changes in soil stiffness
for relatively few load cycles (N < 200). Since no reliable under long-term stress application.
model presently exists, design requirements regarding the Instrumented field tests on full-scale ‘rigid’ monopiles,
pile’s rotation=lateral displacement behavior under mono- combined with a more extensive program of testing on
tonic (static) extreme load are used as a substitute. Areas model-scale piles installed in different soil conditions, con-
warranting further in-depth research are the effects on sidering changing load characteristics (amplitude, frequency,
monopile behavior of: and direction) and subjected to high numbers of load cycles
. Soil properties; e.g., relative density, over-consolidation of more than 106 are warranted. Such studies would provide
ratio, and the relative significance of the at-rest earth valuable information for the validation of current and
pressure coefficient (K0) and the soil stiffness, including improved design methods=theories for offshore monopiles
their variations with depth; and the calibration of pertinent numerical models. They
20 M. Arshad & B. C. O’Kelly

would also be helpful in generating computer code for Engineering 138(3): 364–75. doi:10.1061=(asce)gt.1943-
numerical simulations of more realistic conditions, especially 5606.0000592
in-situ soil conditions and in-service loading characteristics. Bisoi, S. and S. Haldar. 2014. Dynamic analysis of offshore wind tur-
bine in clay considering soil–monopile–tower interaction. Soil
Dynamics and Earthquake Engineering 63: 19–35. doi:10.1016=
Acknowledgment j.soildyn.2014.03.006
Blanco, M. I. 2009. The economics of wind energy. Renewable and Sus-
The first author gratefully acknowledges a Postgraduate tainable Energy Reviews 13(6–7): 1372–82.
Research Scholarship Award from Trinity College Dublin. Broms, B. 1964. Lateral resistance of piles in cohesionless soils. Soil
The writers also thank the reviewers for many helpful Mechanics and Foundation Engineering Division, ASCE 90(3):
comments. 123–56.
Budhu, M. 2011. Soil Mechanics and Foundations. New York: John
Wiley & Sons.
References Byrne, B. W. and G. T. Houlsby. 2003. Foundations for offshore wind
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

Abdel-Rahman, K. and M. Achmus. 2011. Behavior of foundation turbines. Philosophical Transactions of the Royal Society of London
piles for offshore wind energy plants under axial cyclic loading. 361(1813): 2909–30.
Proceedings of the SIMULIA Customer Conference, Barcelona, Carswell, W., S. R. Arwade, D. J. DeGroot, and M. A. Lackner. 2015.
Spain, Dassault Systèmes Press, Paris, France, 17–19th May Soil–structure reliability of offshore wind turbine monopile foun-
2011, 331–41. dations. Wind Energy 18(3): 483–98. doi:10.1002=we.1710
Achmus, M. 2010. Design of axially and laterally loaded piles for the Chai, J.-C. and N. Miura. 2002. Traffic-load-induced permanent defor-
support of offshore wind energy converters. Proceedings of the mation of road on soft subsoil. Journal of Geotechnical and Geoen-
Indian Geotechnical Conference GEOtrendz–2010, Mumbai, vironmental Engineering 128(11): 907–16. doi:10.1061=(asce)1090-
India, December 16–18th, 2010, 92–102. 0241(2002)128:11(907)
Achmus, M., Y.-S. Kuo, and K. Abdel-Rahman. 2009. Behaviour of Chakrabarti, S. K. 2005. Handbook of Offshore Engineering. London,
monopile foundations under cyclic lateral load. Computer and UK: Elsevier.
Geotechnics 36(5): 725–35. doi:10.1016=j.compgeo.2008.12.003 Das, B. M. 2008. Advanced Soil Mechanics. New York: CRC Press.
API (American Petroleum Institute). 2010. Recommended practice for Davisson, M. T. 1970. Lateral Load Capacity of Piles, Vol. 333, 104–12.
planning, designing and constructing fixed offshore platforms— Washington, DC: Highway Research Record.
Working stress design. API RP 2A-WSD (R2010), 22nd ed., De Vries, W. E. 2007. Project UpWind WP4 deliverable D4.2.1: Assess-
API Publishing Services, Washington, DC, USA. ment of bottom-mounted support structure types with conven-
Arshad, M. and B. C. O’Kelly. 2013. Offshore wind-turbine structures: tional design stiffness and installation techniques for typical
A review. Proceedings of the ICE, Energy 166(4): 139–52. deep water sites. http://www.upwind.eu/Publications/~/media/
doi:10.1680=ener.12.00019 UpWind/Documents/Publications/4%20-%20Offshore%20Foun-
Arshad, M. and B. C. O’Kelly. 2014. Development of a rig to study dations/UpwindWP4D421%20Assessment%20of%20bottom-
model pile behaviour under repeating lateral loads. International mounted%20support%20structure%20types.ashx (accessed March
Journal of Physical Modelling in Geotechnics 14(3): 54–67. 15th, 2015).
doi:10.1680=ijpmg.13.00015 De Vries, W. E. and J. van der Tempel. 2007. Quick monopile design.
Atkinson, J. H. and G. Sallfors. 1991. Experimental determination of Proceedings of the European Offshore Wind Conference & Exhi-
stress–strain–time characteristics in laboratory and in-situ tests. bition, Berlin, Germany, December 4–6th, 2007.
Proceedings of the 10th European Conference on Soil Mechanics DIN (Deutsches Institut fur Normung). 2005. Baugrund Sicherheits-
and Foundation Engineering, Florence, Italy. Balkema: Rotter- nachweise im Erd und Grundbau. DIN 1054: 2005. DIN: Berlin,
dam, The Netherlands, May 26–30th, 1991, Vol. 3, 915–56. Germany. http://www.baunormenlexikon.de/Normen/DIN/
Barksdale, R. D. 1972. Laboratory evaluation of rutting in base course DIN%201054/1b69b621-74a0-4c33-8b29-a82a24afd4d4 (accessed
materials. Proceedings of the 3rd International Conference on the March 15th, 2015).
Structural Design of Asphalt Pavement, London, UK, September DNV (Det Norske Veritas). 2011. Design of offshore wind turbine
11–15th, 1972, Vol. 1, 161–74. structures. DNV-OS-J101, DNV, Oslo, Norway.
Bartolomey, A. A. 1977. Experimental analysis of pile groups under lat- Dobry, R., E. Vicenti, M. J. O’Rourke, and J. M. Roesset. 1982. Hori-
eral loads. Proceedings of the 9th International Conference on Soil zontal stiffness and damping of single piles. Journal of Geotechni-
Mechanics and Foundation Engineering, Tokyo, Japan, July 10– cal Engineering Division, ASCE 108(3): 439–59.
15th, 1977, 187–88. Dührkop, J. 2009. On the influence of expanders and cyclic loads on the
Basack, S. and S. Dey. 2012. Influence of relative pile-soil stiffness and deformation behavior of lateral stressed piles in sand. PhD Thesis,
load eccentricity on single pile response in sand under lateral cyclic Hamburg University of Technology.
loading. Geotechnical and Geological Engineering 30(4): 737–51. Dyson, G. J. and M. F. Randolph. 2001. Monotonic lateral loading of
doi:10.1007=s10706-011-9490-1 piles in calcareous sand. Journal of Geotechnical and Geoenviron-
Basack, S. and S. Sen. 2014. Numerical solution of single piles mental Engineering 127(4): 346–52. doi:10.1061=(asce)1090-
subjected to pure torsion. Journal of Geotechnical and Geoenviron- 0241(2001)127:4(346)
mental Engineering 140(1): 74–90. doi:10.1061=(asce)gt.1943-5606. EWEA (European Wind Energy Association). 2011. Design limits and
0000964 solutions for very large wind turbines. UpWind project. http://
Batchelor, G. K. 1967. An Introduction to Fluid Dynamics. Cambridge, www.ewea.org/fileadmin/ewea_documents/documents/upwind/
UK: Cambridge University Press. 21895_UpWind_Report_low_web.pdf (accessed March, 15th 2015).
Bhattacharya, S., J. A. Cox, D. Lombardi, and D. M. Wood. 2012. Fischer, T. 2011. Executive summary – Upwind project WP4: Offshore
Dynamics of offshore wind turbines supported on two founda- foundations and support structures. http://www.upwind.eu/Pub-
tions. Proceedings of the ICE, Geotechnical Engineering 166(2): lications/~/media/UpWind/Documents/Publications/4%20-
159–69. doi:10.1680=geng.11.00015 %20Offshore%20Foundations/WP4_Executive_Summary_Fina-
Bienen, B., J. Dührkop, J. Grabe, M. F. Randolph, and D. J. White. l.ashx (accessed March, 15th 2015).
2012. Response of piles with wings to monotonic and cyclic lateral Gavin, K. G. and O’Kelly, B. C. 2007. Effect of friction fatigue
loading in sand. Journal of Geotechnical and Geoenvironmental on pile capacity in dense sand. Journal of Geotechnical and
Analysis and Design of Offshore Monopile Foundations 21
Geoenvironmental Engineering 133(1): 63–71. doi:10.1061= development. Technical Report NREL=TP-500–38060, National
(asce)1090-0241(2007)133:1(63) Renewable Energy Laboratory (NREL), Colorado, USA.
GL (Germanischer Lloyd). 2005. Guideline for the Certification of Off- Journée, J. M. J. and W. W. Massie. 2001. Offshore Hydrodynamics, 1st
shore Wind Turbines. Hamburg, Germany: GL. ed. Delft, The Netherlands: Delft University of Technology.
Goryunov, B. F. 1973. Analysis of piles subjected to the combined Karg, C. 2007. Modelling of strain accumulation due to low level vibrations
action of vertical and horizontal loads (discussion). Soil Mechanics in granular soils. PhD Thesis. Ghent University, Ghent, Belgium.
and Foundation Engineering 10(1): 10–13. doi:10.1007=bf01706631 Karthigeyan, S., V. V. G. S. T. Ramakrishna, and K. Rajagopal. 2006.
Guo, W. D. 2006. On limiting force profile, slip depth and response of Influence of vertical load on the lateral response of piles in sand.
lateral piles. Computer and Geotechnics 33(1): 47–67. doi:10.1016= Computers and Geotechnics 33(2): 121–31. doi:10.1016=j.comp-
j.compgeo.2006.02.001 geo.2005.12.002
Guo, W. D. 2013. Pu is subscripted-based solutions for slope stabilizing Karthigeyan, S., V. V. G. S. T. Ramakrishna, and K. Rajagopal. 2007.
piles. International Journal of Geomechanics 13(3): 292–310. Numerical investigation of the effect of vertical load on the lateral
doi:10.1061=(asce)gm.1943-5622.0000201 response of piles. Journal of Geotechnical and Geoenvironmental
Haiderali, A., U. Cilingir, and G. Madabhushi. 2013. Lateral and axial Engineering 133(5): 512–21. doi:10.1061=(asce)1090-0241(2007)
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

capacity of monopiles for offshore wind turbines. Indian Geotech- 133:5(512)


nical Journal 43(3): 181–94. doi:10.1007=s40098-013-0056-4 Klinkvort, R. T. and O. Hededal. 2013. Lateral response of monopile
Haritos, N. 2007. Introduction to the analysis and design of offshore supporting an offshore wind turbine. Proceedings of the ICE, Geo-
structures – An overview. Electronic Journal of Structural technical Engineering 166(2): 147–58. doi:10.1680=geng.12.00033
Engineering 7(Special Issue: Loading on Structures): 55–65. Kopp, D. R. 2010. Foundations for an offshore wind turbine. MSc
Hornych, P., J. F. Corte, and J. L. Paute. 1993. Etude des déformations Thesis, Massachusetts Institute of Technology.
permanentes sous chargements répétés de trois graves non traitées. Kuo, Y.-S., M. Achmus, and K. Abdel-Rahman. 2012. Minimum
Bulletin de Liaison des Laboratoires des Ponts et Chaussées, Presse embedded length of cyclic horizontally loaded monopiles. Journal
de l’ENPC, Paris, France 184: 45–55. of Geotechnical and Geoenvironmental Engineering 138(3): 357–63.
Huurman, M. 1996. Development of traffic induced permanent strains doi:10.1061=(asce)gt.1943-5606.0000602
in concrete block pavements. Heron 41(1): 29–52. LeBlanc, C. 2009. Design of offshore wind turbine support structures:
IEC (International Electrotechnical Commission). 2009. Wind turbines Selected topics in the field of geotechnical engineering. PhD The-
– Part 3: Design requirements for offshore wind turbines. IEC sis, Aalborg University, Aalborg, Denmark.
61400–3: 2009, International Electrotechnical Commission, Gen- LeBlanc, C., G. T. Houlsby, and B. W. Byrne. 2010. Response of stiff
eva, Switzerland. piles in sand to long-term cyclic lateral loading. Géotechnique
Igoe, D., K. Gavin, and B. O’Kelly. 2010. Field tests using an instru- 60(2): 79–90. doi:10.1680=geot.7.00196
mented model pipe pile in sand. Proceedings of the Seventh Inter- Lekarp, F. and A. Dawson. 1998. Modelling permanent deformation
national Conference on Physical Modelling in Geotechnics, behaviour of unbound granular materials. Construction and Build-
Zurich, Switzerland, June 28th–July 1st, 2010, ed. S. Springman, ing Materials 12(1): 9–18. doi:10.1016=s0950-0618(97)00078-0
J. Laue, and L. Seward, Vol. 2, 775–780. Leiden, The Netherlands: Lentz, R. W. and G. Y. Baladi. 1981. Constitutive equation for perma-
CRC Press. nent strain of sand subjected to cyclic loading. Transportation
Igoe, D., K. Gavin and B. O’Kelly. 2013. An investigation into the use Research Record 810: 50–54.
of push-in pile foundations by the offshore wind sector. Inter- Lesny, K. and J. Wiemann. 2005. Design aspects of monopiles in Ger-
national Journal of Environmental Studies 70(5): 777–91. man offshore wind farms. Proceedings of the 1st International
doi:10.1080=00207233.2013.798496 Symposium on Frontiers in Offshore Geotechnics. Perth, Austra-
Igoe, D. J. P., K. G. Gavin, B. C. O’Kelly, and B. Byrne. 2013. The use lia, September 19–21st, 2005, ed. S. Gourvenec and M. Cassidy,
of in-situ site investigation techniques for the axial design of off- 383–89. Leiden, The Netherlands: Balkema.
shore piles. Proceedings of the 4th International Conference on Li, D. and E. Selig. 1996. Cumulative plastic deformation for fine-
Geotechnical and Geophysical Site Characterization, Pernam- grained subgrade soils. Journal of Geotechnical Engineering, ASCE
buco, Brazil, September 18–21st, 2012, ed. R. Q. Coutinho and 122(12): 1006–13. doi:10.1061=(asce)0733-9410(1996)122:12(1006)
P. W. Mayne. Vol. 2, 1123–29. CRC Press. Li, Z., S. K. Haigh, and M. D. Bolton. 2010. Centrifuge modelling of
Irvine, J. H., P. G. Allan, B. G. Clarke, and J. R. Peng. 2003. Improv- mono-pile under cyclic lateral loads. Proceedings of the 7th Inter-
ing the lateral stability of monopile foundations. Proceedings of national Conference on Physical Modelling in Geotechnics, Zur-
the BGA International Conference on Foundations: Innovations, ich, Switzerland, 28th June–1st July 2010, ed. S. Springman, J.
Observations, Design and Practice, Dundee, UK, September 2– Laue, and L. Seward, Vol. 2, 965–70. Leiden, The Netherlands:
5th, 2003, ed. T. A. Newson, 371–80. London: Thomas Telford. CRC Press.
Ismael, N. F. 1990. Behavior of laterally loaded bored piles in cemented Lin, S.-S. and J.-C. Liao. 1999. Permanent strains of piles in sand due
sands. Journal of Geotechnical Engineering, ASCE 116(11): 1678– to cyclic lateral loads. Journal of Geotechnical and Geoenvironmen-
99. doi:10.1061=(asce)0733-9410(1990)116:11(1678) tal Engineering 125(9): 798–802. doi:10.1061=(asce)1090-
Jaimes, O. G. 2010. Design concepts for offshore wind turbines: 0241(1999)125:9(798)
A technical and economic study on the trade-off between stall Little, R. L. and J.-L. Briaud. 1988. Full scale cyclic lateral load tests on
and pitch controlled systems. MSc Thesis, Delft University of six single piles in sand. Miscellaneous paper GL-88–27, Geotech-
Technology. nical Division, Civil Engineering Department, Texas A&M Uni-
Jain, N. K., G. Ranjan, and G. Ramasamy. 1987. Effect of vertical load versity, College Station, TX, USA.
on flexural behaviour of piles. Geotechnical Engineering: Journal of Lombardi, D., S. Bhattacharya, and D. M. Wood. 2013. Dynamic soil–
the Southeast Asian Geotechnical Society 18(2): 185–204. structure interaction of monopile supported wind turbines in
Jang, J. J. and G. J. Shinn. 1999. Analysis of maximum wind force for cohesive soil. Soil Dynamics and Earthquake Engineering 49:
offshore structure design. Journal of Marine Science and Tech- 165–80. doi:10.1016=j.soildyn.2013.01.015
nology 7(1): 43–51. Long, J. H. and G. Vanneste. 1994. Effects of cyclic lateral loads on
Jonkman, J., S. Butterfield, W. Musial, and G. Scott. 2009. Definition piles in sand. Journal of Geotechnical Engineering, ASCE 120(1):
of a 5-MW reference wind turbine for offshore system 225–44. doi:10.1061=(asce)0733-9410(1994)120:1(225)
22 M. Arshad & B. C. O’Kelly

Malhotra, S. 2011. Design and construction considerations for offshore Peng, J., B. G. Clarke, and M. Rouainia. 2011. Increasing the resistance
wind turbine foundations in North America. Proceedings Geo- of piles subject to cyclic lateral loading. Journal of Geotechnical
Florida 2010: Advances in Analysis, Modeling and Design, West and Geoenvironmental Engineering 137(10): 977–82. doi:10.1061=
Palm Beach, Florida, USA, February 20–24th, 2010, ed. D. O. (asce)gt.1943-5606.0000504
Fratta, A. J. Puppala, and B. Muhunthan, Vol. 2, 1533–42, GSP Peralta, P. and M. Achmus. 2010. An experimental investigation of
199. Red Hook, NY, USA: Curran Associates, Inc. piles in sand subjected to lateral cyclic loads. Proceedings of the
Matlock, H. 1970. Correlation for design of laterally loaded piles in soft 7th International Conference on Physical Modelling in Geotech-
clay. Proceedings of the 2nd Offshore Technology Conference, 22– nics, Zurich, Switzerland, 28th June–1st July 2010, ed. S. Spring-
24th April 1970, Houston, TX, USA. Paper Number 1204. Vol. 1, man, J. Laue, and L. Seward, Vol. 2, 985–90. Leiden, The
77–94. Netherlands: CRC Press.
Matlock, H., and L. C. Reese 1960. Generalized solutions for laterally Poulos, H. G. 1988. Cyclic stability diagram for axially loaded piles.
loaded piles. Journal of the Soil Mechanics and Foundations Journal of Geotechnical Engineering, ASCE, 114(8): 877–95.
Division ASCE 86(5): 63–94. doi:10.1061=(asce)0733-9410(1988)114:8(877)
Moayed, R. Z., I. Mehdipour, and A. Judi. 2012. Undrained lateral Rahim, A. and R. F. Stevens. 2013. Design procedures for marine
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

behavior of short pile under combination of axial, lateral and renewable energy foundations. Proceedings of the 1st Marine
moment loading in clayey soils. Kuwait Journal of Science and Energy Technology Symposium, Washington, DC, April 10–
Engineering 39(1B): 59–78. 11th, 2013, 10.
Morison, J. R., J. W. Johnson, and S. A. Schaff. 1950. The forces Ramasamy, G. 1974. Flexural behaviour of axially and laterally loaded
exerted by surface waves on piles. Journal of Petroleum Technology individual piles and groups of piles. PhD Thesis, Indian Institute
2(5): 149–54. doi:10.2118=950149-g of Science, Bangalore.
Naughton, P. J. and B. C. O’Kelly. 2004. The induced anisotropy of Randolph, M. F. 1981. The response of flexible piles to lateral loading.
Leighton Buzzard sand. Proceedings Advances in Geotechnical Géotechnique 31(2): 247–59. doi:10.1680=geot.1981.31.2.247
Engineering: The Skempton Conference, London, UK, March Rani, S. and A. Prashant. 2014. Estimation of the linear spring con-
28th–31st, 2004, ed. R. J. Jardine, D. M. Potts, and K. G. Higgins, stant for a laterally loaded monopile embedded in nonlinear soil.
Vol. 1, 556–67. London, UK: Thomas Telford. International Journal of Geomechanics. doi:10.1061=
Naughton, P. J. and B. C. O’Kelly. 2005. Yield behavior of sand under (ASCE)GM.1943-5622.0000441. Online Publication Date: 29th
generalized stress conditions. Proceedings of the 16th International August 2014.
Conference on Soil Mechanics and Geotechnical Engineering, Osaka, Reese, L. C, W. R. Cox, and F. D. Koop. (1974). Analysis of laterally
Japan, September 12–16th, 2005. IOS Press. Vol. 2, 555–58. loaded piles in sand. Proceedings of the 6th Annual Offshore Tech-
Nicolai, G. and L. B. Ibsen. 2014. Small-scale testing of cyclic laterally nology Conference, Houston, Texas, May 6–8th, 1974, 473–84.
loaded monopiles in dense saturated sand. Journal of Ocean and Reese, L. C, W. R. Cox, and F. D. Koop. (1975). Field testing
Wind Energy 1(4): 240–45. and analysis of laterally loaded piles in stiff clay. Procedings
Niemunis, A., T. Wichtmann, and Th. Triantafyllidis. 2005. A high- of the 7th Annual Offshore Technology Conference. Paper No.
cycle accumulation model for sand. Computers and Geotechnics 2312, Houston, Texas, May 5–8th, 1975, 671–690. doi: 10.4043/
32(4): 245–63. doi:10.1016=j.compgeo.2005.03.002 2312-MS.
O’Kelly, B. C. 2005. Consolidation anisotropy of some natural soft Richwien, W., K. Lesny, and J. Wiemann. 2002. Bau- und umwelttech-
soils. Proceedings of the International Conference on Problematic nische Aspekte von Off-shore Windenergieanlagen (Construction
Soils, Famagusta, North Cyprus, May 25–27th, 2005, ed. H. Bilsel and environmental aspects of offshore wind turbines). Gigawind
and N. Zalihe, Vol. 3, 1183–92. North Cyprus: Eastern Mediterra- Annual Report 2001, Chapter 6: Support structure – Foundation,
nean University Press. 36–46. Leibniz Universität Hannover, Germany.
O’Kelly, B. C. 2006. Compression and consolidation anisotropy of Şahin, A. D. 2004. Progress and recent trends in wind energy. Progress
some soft soils. Geotechnical and Geological Engineering 24(6): in Energy and Combustion Science 30(5): 501–43. doi:10.1016=
1715–28. doi:10.1007=s10706-005-5760-0 j.pecs.2004.04.001
O’Kelly, B. C. and P. J. Naughton. 2005a. Development of a new hol- Sorochan, E. A. and V. I. Bykov. 1976. Performance of groups of
low cylinder apparatus for stress path measurements over a wide cast-in place piles subjected to horizontal loading. Soil Mechanics
strain range. Geotechnical Testing Journal 28(4): 345–54. and Foundation Engineering 13(3): 157–61. doi:10.1007=
doi:10.1520=gtj12252 bf01705310
O’Kelly, B. C. and P. J. Naughton. 2005b. Engineering properties of St. Denis, M. and W. J. Pierson. 1953. On the motions of ships in con-
wet-pluviated hollow cylindrical specimens. Geotechnical Testing fused seas. Society of Naval Architects and Marine Engineers
Journal 28(6): 570–76. doi:10.1520=gtj12325 Transactions 61: 280–354.
O’Kelly, B. C. and P. J. Naughton. 2008. Local measurements of the Sweere, G. T. H. 1990. Unbound granular bases for roads. PhD Thesis,
polar deformation response in a hollow cylinder apparatus. Geo- Delft University of Technology.
mechanics and Geoengineering 3(4): 217–29. doi:10.1080= Terzaghi, K. 1955. Evaluation of coefficients of subgrade reaction.
17486020802400981 Géotechnique 5(4): 297–326.
O’Kelly, B. C. and P. J. Naughton. 2009. Study of the yielding of sand Timmerman, D. H. and T. H. Wu. 1969. Behavior of dry sands under
under generalized stress conditions using a versatile hollow cylin- cyclic loading. Journal of the Soil Mechanics and Foundations
der torsional apparatus. Mechanics of Materials 41(3): 187–98. Division ASCE 95(4): 1097–114.
doi:10.1016=j.mechmat.2008.11.002 Tomlinson, M. J. 2001. Foundation Design and Construction, 7th edn.
Pappin, J. W. 1979. Characteristics of granular material for pavement Harlow, England: Pearson Education.
analysis. PhD Thesis, University of Nottingham. Tong, W. 2010. Wind Power Generation and Wind Turbine Design.
Paute, J.-L., P. Hornych, and J. P. Benaben. 1996. Repeated load triax- Southampton, UK: WIT Press.
ial testing of granular materials in the French Network of Labora- van der Tempel, J. 2006. Design of support structures for offshore wind
tories des Ponts et Chaussées. Flexible Pavements: Proceedings of turbines. PhD Thesis, Delft University of Technology.
the European Symposium Euroflex, Lisbon, Portugal, September Verdure, L., J. Garnier, and D. Levacher. 2003. Lateral cyclic loading
20–22nd, 1993, ed. A. G. Corriea, 53–64. Rotterdam, The of single piles in sand. International Journal of Physical Modelling
Netherlands: Balkema. in Geotechnics 3(3): 17–28.
Analysis and Design of Offshore Monopile Foundations 23
Wang, Q. and P. V. Lade. 2001. Shear banding in true triaxial tests and Appendix
its effect on failure in sand. Journal of Engineering Mechanics
127(8): 754–61. doi:10.1061=(asce)0733-9399(2001)127:8(754)
Werkmeister, S. 2003. Permanent deformation behaviour of unbound
granular materials in pavement constructions. PhD Thesis, Tech-
nical University Dresden.
Wichtmann, T., A. Niemunis, and Th. Triantafyllidis. 2009. Validation
and calibration of a high-cycle accumulation model based on cyc-
lic triaxial tests on eight sands. Soils and Foundations 49(5): 711–
28. doi:10.3208=sandf.49.711
Wichtmann, T., H. A. Rondón, A. Niemunis, Th. Triantafyllidis, and
A. Lizcano. 2010. Prediction of permanent deformations in pave-
ments using a high-cycle accumulation model. Journal of Geotech-
nical and Geoenvironmental Engineering 136(5): 728–40.
Downloaded by [University of Engineering & Technology Lahore] at 01:19 03 February 2016

doi:10.1061=(asce)gt.1943-5606.0000275
Yasin, S. J. M. and F. Tatsuoka. 2000. Stress history-dependent defor-
mation characteristics of dense sand in plane strain. Soils and
Foundations 40(2): 77–98. doi:10.3208=sandf.40.2_77
Zhu, B., B. W. Byrne, and G. T. Houlsby. 2013. Long-term lateral cyc-
lic response of suction caisson foundations in sand. Journal of
Geotechnical and Geoenvironmental Engineering 139(1): 73–83.
doi:10.1061=(asce)gt.1943-5606.0000738
Zhukov, N. V. and I. L. Balov. 1978. Investigation of the effect of a
vertical surcharge on horizontal displacements and resistance of Fig. A1. Pertinent stiffness profiles for the medium dense sand
pile columns to horizontal loads. Soil Mechanics and Foundation deposit considered in the OWT monopile foundation design
Engineering 15(1): 16–22. doi:10.1007=bf02145324 example.
Model Studies on Monopile Behavior under Long-Term
Repeated Lateral Loading
Muhammad Arshad, Ph.D.1; and Brendan C. O’Kelly, Ph.D., C.Eng.2
Downloaded from ascelibrary.org by University of Engineering and Technology, Lahore on 05/03/16. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Monopiles are the most commonly used foundation type for offshore wind turbine (OWT) structures and are characterized by rela-
tively large geometric dimensions, compared with offshore pile foundations typically used in the oil and gas industries. To date, there are no
established technical guidelines tailored for the design and analysis of OWT monopiles. This paper first identifies various intrinsic drawbacks
involved with the existing design and analysis methodologies as applied to OWT monopiles. Next, a comprehensive experimental program of
1g repeated lateral load tests, performed on a scaled rigid monopile installed in dry sand beds, is presented to investigate its behavior under var-
ious loading scenarios. The experimental results provide insights into the various blurry issues in the existing literature related to monopile
behavior under long-term repeated lateral loading. Lateral soil resistance profiles were determined from the measured pile bending strain data
and found to be markedly dependent on the degree of the polynomial function used for curve-fitting of the bending strain data. Finally, an ex-
perimental model is presented for estimation of the pile’s accumulated rotation, which takes into account various basic characteristics of the
applied lateral load cycles. DOI: 10.1061/(ASCE)GM.1943-5622.0000679. © 2016 American Society of Civil Engineers.
Author keywords: Cyclic loading; Footings; Foundations; Lateral loading; Model tests; Offshore engineering.

Introduction and DNV (2011) do not provide the means of calculating the accu-
mulated pile displacement (rotation), and they always imply stiffness
For offshore wind turbine (OWT) structures, the horizontal/lateral degradation of the soil–pile foundation system as a result of long-
loads (and resulting moments about the foundation) are generally term repeated lateral loading, irrespective of soil type or soil state.
large in proportion to the gravitational loads (Byrne and Houlsby However, this is contrary to experimental observations documented
2003). The foundation design is dominated by considerations of the by LeBlanc et al. (2010), Bhattacharya et al. (2011), Bhattacharya
dynamic response and fatigue under working loads, rather than ulti- and Adhikari (2011), Cuellar et al. (2012), and Arshad and O’Kelly
mate loading conditions. Monopile foundations for OWTs typically (2014, 2016a). Further, disputing interpretations regarding stiffness
have an embedment length–to–outer diameter ratio (L/D) of less determination are not uncommon in the literature, possibly due to
than 10, categorizing them as rigid piles (Tomlinson 2001; Peng et discrepancies in the terminology used by the different researchers.
al. 2011; DNV 2011). Hence, they are more prone to rotation from For instance, the term stiffness has been reported in some literature
their initial vertical alignment (as compared to deflection) under the without specifying whether it relates to the absolute secant, tangent,
action of the repeating lateral loads and moments (Haiderali et al. or cyclic stiffness scenario. The following are also common open
questions in implementing the p–y method for the analysis of OWT
2014). Current pile design methodology based on p–y curves, as pre-
monopile foundations: (1) influence of vertical pile load on the lateral
sented in the American Petroleum Institute (API 2010) and Det
response of the soil–pile system, (2) diameter effect on the initial
Norske Veritas (DNV 2011) codes, has gained widespread recogni-
stiffness of the p–y curves, (3) choice of horizontal earth pressure
tion on account of the low failure rate of in-service piles, evidenced
coefficient, (4) pilehead fixity, (5) pile installation effects, and (6)
over many decades (Arshad and O’Kelly 2013). However, the API
shearing force developed at the pile base.
(2010) and DNV (2011) guidelines for monopile design mainly rely The inadequacy of the current p–y methodology for predicting
on methods developed using empirical data obtained for long flexi- the response of rigid monopiles under repeated lateral loading has
ble piles for which bending is generally critical. Hence, design crite- been reported by many researchers (e.g., Haiderali and Madabhushi
ria and analyses for rigid pile scenarios cast considerable doubt on 2013; Lau et al. 2014; Arshad and O’Kelly 2016b; O’Kelly and
the application of such methods for OWT monopiles. Further, the Arshad 2016), which has led to the development of new models
p–y curves for repeated lateral loading presented in the API (2010) for obtaining better predictions of the strain accumulation in the
soil surrounding the monopile, with due consideration of the soil
1
Assistant Professor, Dept. of Geological Engineering, Univ. of –pile interaction. However, the proposed models also have in-
Engineering and Technology, Lahore 54890, Pakistan; formerly, Ph.D. herent limitations. For instance, the models proposed by Little
candidate, Dept. of Civil, Structural and Environmental Engineering, and Briaud (1988), Long and Vanneste (1994), and Lin and
Trinity College Dublin, Dublin, Dublin 2, Ireland (corresponding author). Liao (1999) are based on limited numbers of lateral load cycles
E-mail: arshadm@tcd.ie
2
N (<100) applied to long flexible piles installed in sandy soil,
Associate Professor, Dept. of Civil, Structural and Environmental with these models always suggesting degradation of the soil–pile
Engineering, Trinity College Dublin, Dublin, Dublin 2, Ireland. E-mail:
system under the repeating loads. The model proposed by
bokelly@tcd.ie
Note. This manuscript was submitted on December 30, 2014; approved Klinkvort and Hededal (2013) is based on a maximum of 500
on February 23, 2016; published online on April 26, 2016. Discussion pe- load cycles performed at a single value of loading frequency
riod open until September 26, 2016; separate discussions must be submit- and for a particular soil density. Although based on many thou-
ted for individual papers. This paper is part of the International Journal sands of load cycles, the model proposed by LeBlanc et al.
of Geomechanics, © ASCE, ISSN 1532-3641. (2010) appears to be unrealistic because it always predicts zero

© ASCE 04016040-1 Int. J. Geomech.

Int. J. Geomech., 04016040


displacement (rotation) of the monopile under two-way balanced for OWT monopile foundations have been reported by Lai (1989),
repeat loading. Muir Wood et al. (2002), LeBlanc et al. (2010), Bhattacharya et al.
This paper presents an experimental study performed on a scaled (2011), and Cuellar et al. (2012).
rigid monopile subjected to long-term repeated lateral loading. The
experimental results are discussed in relation to various controver-
sial issues relating to the long-term performance of monopile foun- Testing Facility, Model Pile, and Instrumentation
dations under repeated lateral loading. A new experimental model
is proposed to estimate the accumulated rotation of the monopile, For the present investigation, the Trinity College Dublin electro-
adequately incorporating many thousands of load cycles of different mechanical loading system, which can apply many thousands of lat-
load amplitude and cyclic character. Different shapes of the lateral eral load cycles of sinusoidal waveform shape to the head (top) of a
Downloaded from ascelibrary.org by University of Engineering and Technology, Lahore on 05/03/16. Copyright ASCE. For personal use only; all rights reserved.

soil resistance (LSR) profile are evaluated on the basis of bending scaled monopile embedded in a sand bed, was used (Fig. 1).
strain data obtained along the embedded length of the instrumented Different loading schemes can be applied to the pilehead, with full
monopile. control provided over the loading direction, load amplitude, and fre-
quency. Further details on the working mechanisms and capabilities
of the loading system, the instrumentation and data-acquisition sys-
Scaling Issues for Model Testing tem used to measure the applied lateral loads and the displacement
(rotation) response at the pilehead (Fig. 2), the sample (sand bed)
An important task for reduced-scale model testing is determining preparation technique, and the method of pile installation (nondis-
how to generalize the results for prototype cases. Differences in a placement) are presented in Arshad and O’Kelly (2014). Some
number of aspects, such as the soil stresses developed and the rela- properties of the test sand are given in Table 1.
tive soil–structure size, can create dissimilarities in the soil–structure The 540-mm-long pile was manufactured from brass tubing
interaction and response between the model and prototype, and so
(Young’s modulus value of 96 GPa) with an outer diameter of 53.0
often invalidate simple extrapolation of the 1g (i.e., normal gravity
and a wall thickness of 0.8 mm, which produced a bending stiffness
condition) data to large-scale cases. Nevertheless, scaled model test-
(EI) value of 4.33 kN·m2. The lower end of the pile was closed by
ing can be useful if it preserves geometric, kinematic, or dynamic
using 3-mm-thick brass plate to represent a fully plugged tubular pile.
similarity to the prototype. Among other technical difficulties, the
The dimensions of the model pile typically represent a field monopile
direct scaling of the soil particle (grain) dimensions is particularly
(made of steel) for an OWT foundation system at 1/100th scale.
problematic because it might introduce undesirable forces into play
The outer wall surface of the model pile was instrumented with
unless a certain minimum ratio is maintained between the mean
8-strain gauges of type TML–PL–10–11 (manufactured by Tokyo
grain size (d50) and a characteristic dimension (pile diameter in the
present investigation) of the model (Sedran et al. 2001; Verdure Sokki Kenkyujo, Tokyo), which were equi-spaced over its embed-
et al. 2003). As reported by Cuellar (2011), model tests performed ment length of 360 mm. The center of the lowermost gauge was
with prototype sand might show disproportionate shear bands and located at a distance of 10 mm above the pile toe level, with the
the mobilization of higher ultimate loads, although he goes on to uppermost gauge located just beneath the finished sand bed surface
point out that “several studies have shown that the influence of level in the 0.95 m diameter by 0.6-m-deep steel tank (Item 5 in
these shear-zones on the overall response of the foundation may Fig. 1). The dimensions of the steel tank were chosen to ensure that
be neglected if the applied loads lay well below the failure limit” the failure wedge developed around the pile for the static and
(Hettler 1981) “or if the ratio of foundation diameter to grain size repeated lateral loading scenarios investigated did not extend to the
(D/d50) is greater than 30 to 60” (Ovesen 1979; Franke and Muth tank boundaries. With a tank diameter–to–pile diameter ratio of
1985; Sedran et al. 2001; Verdure et al. 2003). In the present approximately 18, side wall boundary effects were not considered
investigation, a real-size sand was used because both of these con- significant (Davie and Sutherland 1978; Rao et al. 1996). LeBlanc
ditions were reasonably met for the model tests performed. et al. (2010) reported that a soil cushion with a depth of 3–4  D,
Further, the constitutive laws describing the soil load–deformation located below the pile toe, was considered sufficient to absorb the
behavior are stress-level dependent, which implies that for 1g condi- vertical stress field. In our experimental setup, the sand cushion
tions, homologous points of the soil foundations for the model and located beneath the pile toe was 3.4  D in thickness and, further-
prototype might experience different deformational responses. For more, the only vertical loading was caused by the self-weight of the
instance, the load–deformation response of a structure founded on model pile. In the present investigation, the model pile was partially
sand is governed, among other factors, by the magnitude of the mean embedded in dry sand beds prepared at a density of 1577 6 6 kg/m3
in situ effective confining pressure, which is certainly different com- [density index of 70–74% (i.e., dense state)].
paring the model tests under 1g conditions and the prototype (Arshad
and O’Kelly 2014; LeBlanc et al. 2010). For a model test using a shal-
low soil container, the isotropic stress level controlling the mecha- Testing Program
nisms under investigation is low, which leads to higher mobilized fric-
tion angle and low shear modulus values, when compared to its The ultimate static lateral load-carrying capacity (Pu) value of the
equivalent prototype. Many researchers address this issue by perform- model pile (for the given height of 90 mm used as the point of lateral
ing their 1g tests on sand beds prepared at a lower relative density load application above the sand bed surface level) was deduced as
(Kelly et al. 2006; LeBlanc et al. 2010). 140 N. This load value corresponds to a point on the experimental
Similitude relationships in the form of nondimensional parame- load against rotation curve where the pile had rotated from its initial
ters (NDPs) can be applied to relate the results for the model to vertical alignment by 1.5°, producing a lateral displacement at the
those for the prototype, and vice versa (LeBlanc et al. 2010). These sand bed surface level of 7 mm (i.e., 0.13 D). This is in the range
NDPs consider the applied loading, the geometry and material prop- of pile lateral displacements (at the sand bed surface level) of
erties of the pile, and its lateral displacement (rotation) response, between 0.1 and 0.2 D, often used to estimate the Pu value of later-
with due consideration of the pile toe displacement occurring under ally loaded piles (El Sawwaf 2006; Peng et al. 2011; Cuellar et al.
repeated lateral loading. Examples of the derivation of scaling laws 2012).

© ASCE 04016040-2 Int. J. Geomech.

Int. J. Geomech., 04016040


Downloaded from ascelibrary.org by University of Engineering and Technology, Lahore on 05/03/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Schematic diagram of the Trinity College Dublin experimental rig for model pile studies, showing setup for two-way lateral loading of the
pilehead. 1, drive motor; 2, upper-right chain segment; 3, lower-right chain segment; 4, right loading hanger, with weight; 5, steel tank; 6, reaction
frame; 7, right load cell; 8, right spring; 9, articulated arm; 10, sliding node; 11, left loading hanger, with weight; 12, reference support with two hori-
zontal displacement transducers; 13, model pile (adapted from Arshad and O’Kelly 2014, with permission)

Fig. 2. Arrangement of load cells and displacement transducers at the pilehead (adapted from Arshad and O’Kelly 2014, with permission)

Table 1. Properties of Test Sand Different loading schemes, designed on the basis of various loading
directions (balanced and unbalanced two-way and one-way load-
Property Value (mm) ing), load amplitudes, and frequencies, were investigated. The tests
Effective grain sizes, d10, d30, d50, d60 0.16, 0.22, 0.27, 0.31 (loading schemes), listed in Table 2, are identified as follows: load-
Coefficient of uniformity 1.94 ing direction (1w, one way; 2w, two way)/load amplitude (N)/fre-
Coefficient of curvature 1.0 quency (Hz). For example, 1w/0–60/0.25 indicates one-way lateral
Maximum void ratio 0.92 loading having an amplitude of 60 N and a frequency of 0.25 Hz;
Minimum void ratio 0.60 2w/30–60/0.25 indicates two-way lateral loading, 30 N applied in
Density index of sand beds 70–74% one direction and 60 N in the other (i.e., unbalanced) at 0.25 Hz;
Dry sand peak friction angle 39° 1w/15–50/0.25 indicates partial one-way lateral loading, with the
load amplitude fluctuating from 15 to 50 N from the same side, at a
frequency of 0.25 Hz. All of the tests presented in this paper were
A total of 18 lateral load tests, each involving the application of performed at a loading frequency value of 0.25, which corresponds
6,000 lateral load cycles to the pilehead, were performed on the to the typical frequency of the fatigue loading caused by the ocean
model pile embedded in the identically prepared dry sand beds. waves, and if scale effects are considered, it almost corresponds to

© ASCE 04016040-3 Int. J. Geomech.

Int. J. Geomech., 04016040


Table 2. Lateral Load Testing Program on Model Pile

Loading scheme, Test ID Loading direction Left hanger load (N) Right hanger load (N) Frequency (Hz) z b = Pmax/Pu z c = Pmin/Pmax
1w/0–40/0.25 One way 0 40 0.25 0.28 0
1w/12–40/0.25 Partial one way 0 12–40 0.25 0.28 0.3
1w/20–40/0.25 Partial one way 0 20–40 0.25 0.28 0.5
2w/20–40/0.25 Unbalanced two way 20 40 0.25 0.28 −0.5
2w/30–40/0.25 Unbalanced two way 30 40 0.25 0.28 −0.75
2w/40–40/0.25 Balanced two way 40 40 0.25 0.28 −1
1w/0–50/0.25 One way 0 50 0.25 0.35 0
Downloaded from ascelibrary.org by University of Engineering and Technology, Lahore on 05/03/16. Copyright ASCE. For personal use only; all rights reserved.

1w/15–50/0.25 Partial one way 0 15–50 0.25 0.35 0.3


1w/25–50/0.25 Partial one way 0 25–50 0.25 0.35 0.5
2w/25–50/0.25 Unbalanced two way 25 50 0.25 0.35 −0.5
2w/33–50/0.25 Unbalanced two way 33 50 0.25 0.35 −0.66
2w/50–50/0.25 Balanced two way 50 50 0.25 0.35 –1
1w/0–60/0.25 One way 0 60 0.25 0.42 0
1w/18–60/0.25 Partial one way 0 18–60 0.25 0.42 0.3
1w/30–60/0.25 Partial one way 0 30–60 0.25 0.42 0.5
2w/30–60/0.25 Unbalanced two way 30 60 0.25 0.42 −0.5
2w/45–60/0.25 Unbalanced two way 45 60 0.25 0.42 −0.75
2w/60–60/0.25 Balanced two way 60 60 0.25 0.42 −1

the typical energy-rich wind turbulence acting on offshore struc- the soil–pile system under repeated loading, as reported by
tures in the North Sea (LeBlanc 2009). However, the satisfaction of Bhattacharya et al. (2011) and Cuellar et al. (2012).
such scaling conditions for similarity in heterogeneous materials Under repeated lateral loading, Achmus et al. (2009), API (2010),
such as granular soil is not trivial (Dong et al. 2001; Bhattacharya and DNV (2011) suggest that degradation of the soil–pile system
and Adhikari 2011). This situation becomes more complex for sit- occurs, whereas Rosquoët et al. (2007), LeBlanc et al. (2010),
uations such as OWT foundation structures, where one has dynam- Bhattacharya et al. (2011), and Cuellar et al. (2012) have documented
ics, aerodynamics, soil–structure interaction, and fluid flow around that the foundation stiffness actually increases with increasing N, on
the foundations. Other lateral repeating load characteristics consid- account of densification of the soil around the monopile. These con-
ered in this investigation are defined in Eqs. (1) and (2). tradicting interpretations might arise due to discrepancies in the ter-
minology used by different researchers in the various literature. For
Pmax example, in some cases, the generic term stiffness is reported and
zb ¼ (1)
Pu used without specifying whether it refers to initial tangent stiffness,
absolute secant stiffness, or cyclic stiffness values. Cuellar (2011)
Pmin pointed out that the term stiffness has sometimes been incorrectly
zc ¼ (2) used in describing different things, such as (1) the shear modulus of
Pmax the soil, (2) the macromechanical load–displacement relationship for
the foundation system, and (3) the subgrade reaction modulus.
where Pu is the ultimate lateral load-carrying capacity of the pile; During the present experimental investigation, it was observed
Pmax and Pmin are the maximum and minimum lateral loads, respec- (Fig. 3) that the initial few cycles substantially contributed to the
tively, applied to the pilehead during the repeated loading tests. total accumulated rotation of the model pile produced after many
In the present investigation, load amplitudes of 40, 50, and 60 thousands of load cycles had been applied. This observation implies
N (i.e., z b  0.28, 0.35, and 0.42) were considered, with the 40 that the entire experimental curve for all the numbers of load cycles
and 60 N values corresponding to the fatigue and serviceability cannot be described by a single formulation while ensuring reasona-
limit states, respectively, in relation to the pile’s ultimate static lat- ble accuracy. LeBlanc et al. (2010) suggested the following bifur-
eral load-carrying capacity of 140 N (DNV 2011). cating expression can be used to describe the total accumulated
rotation of the monopile:

u N ¼ u 0 þ Du ðNÞ (3)
Rotation Accumulation Model for Monopile under
Cyclic Loading where u N is the accumulated rotation produced for N load cycles;
u 0 is the rotation for the first load cycle; and Du ðNÞ is the rotation
From Fig. 3, it can be seen that the normalized accumulated rotation produced from the second load cycle up to the Nth load cycle. Note
[Du ðNÞ=u 0  increases with the number of load cycles, but at a the value of u 0 is assumed equal to the rotation that would occur in
decreasing rate. This implies that the cyclic stiffness of the soil–pile a monotonic test for a load amplitude equal to the maximum value
system increases with increasing N. Compared with the absolute se- of the load cycle achieved during cyclic loading.
cant stiffness (ratio of the applied load to the absolute displace- For the present investigation, two potentially promising options
ment), the cyclic stiffness might replicate more precisely the actual of curve fitting, namely logarithmic [Eq. (4)] and power [Eq.(5)]
state of the soil–pile system, particularly when its dynamic behavior law approaches, were analyzed by using the experimental data sets.
is under consideration. By contrast, the absolute secant stiffness The logarithmic law technique of curve-fitting to relate the experi-
will always have a decreasing trend, although such degradation has mental model parameters to the load cycle characteristics has been
no physical significance when evaluating the dynamic behavior of used previously; e.g., LeBlanc (2009), LeBlanc et al. (2010),

© ASCE 04016040-4 Int. J. Geomech.

Int. J. Geomech., 04016040


Downloaded from ascelibrary.org by University of Engineering and Technology, Lahore on 05/03/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 3. Curve fitting of normalized accumulated rotation against number of load cycle relationships for various loading schemes

© ASCE 04016040-5 Int. J. Geomech.

Int. J. Geomech., 04016040


Klinkvort and Hededal (2013), and Zhu et al. (2013). However, the
proposed logarithmic law option, having model parameters a1 and
b 1 , is novel in that it uses an additive approach rather than the mul-
tiplicative approach involving different model parameters used in
existing models (e.g., Little and Briaud 1988; Niemunis et al. 2005;
Wichtmann et al. 2010; LeBlanc et al. 2010).
Sample results of regression analysis for the proposed Eqs. (4)
and (5) are shown in Fig. 3, with normalized accumulated rotation
[Du ðNÞ=u 0  plotted against the number of load cycles. From this
figure, the coefficient of determination (R2) values for the logarith-
Downloaded from ascelibrary.org by University of Engineering and Technology, Lahore on 05/03/16. Copyright ASCE. For personal use only; all rights reserved.

mic law approach were relatively higher when compared with the
corresponding values for the power law approach, suggesting that
the logarithmic law captures the accumulation of the pile rotation
with better accuracy.

Du ðNÞ
¼ a1 lnðN Þ þ b 1 (4)
u0

Du ðNÞ
¼ a2 ðNÞ b 2 (5)
u0

where a1 , a2 , b 1 , and b 2 are model parameters.


The power law approach showed reasonable accuracy for the
one-way and partial one-way loading scenarios ( z c ≥ 0 in Fig. 3)
but was found to be markedly inaccurate for the balanced and unbal-
anced two-way loading scenarios ( z c < 0), underestimating the
measured accumulated rotation occurring up to N = 2,000 and over-
estimating this rotation from there up to N = 6,000.
The additive approach of the logarithmic law technique [Eq. (4)]
for the influencing factors ( z c ; z b Þ had the following advantages
over the power law:
1. The power law was found to produce lower R2 values for the
data sets analyzed, and furthermore, a methodical relationship
between a2 and z c or between b 2 and z b could not be
established. Fig. 4. Functions relating (a) a1 with z c and (b) b 1 with z b
2. Calculation errors might be intolerably high with the multipli-
cative approach [Eq. (5)] if either of the a2 and b 2 parameter
values was wrongly estimated.
3. The logarithmic law approach could capture the response for through the origin) exists between b 1 and z b , which is also con-
the initial few cycles, which was found to be substantially de- vincing, because rotation of the monopile cannot be expected for
pendent on z b , with reasonable accuracy. z b = 0 (i.e., b 1 = 0).
Regarding the logarithmic law approach, from regression analy-
sis of the experimental data, the parameter a1 was generalized as a
function of z c [= (Pmin/Pmax)] and the parameter b 1 , as a function Variations in LSR
of z b [= (Pmax/Pu)], as illustrated in Figs. 4(a and b), respectively.
In other words, in applying Eq. (4), the values of a1 and b 1 can be Analytically, the ultimate lateral load-carrying capacity (Pu) of a
estimated from Figs. 4(a and b). These figures realistically capture rigid pile is calculated from the lateral soil resistance (LSR) dis-
the behavior of the monopile subjected to repeated lateral loading tribution profile [denoted by pðzÞ, having unit of force/length]
and show that accumulated rotation always occurs, even for the bal- over its embedded length. The LSR is mobilized in response to
anced two-way loading scenario [i.e., a1 > 0 for z c ¼ 1 in Fig. 4 the lateral displacement of the rigid pile when a lateral load (H) is
(a)], which is a realistic approach. In fact, the direction of the first applied to the pilehead. In general, the LSR can be expressed as
quarter of a two-way load cycle decides the direction of the net the sum of the frontal normal resistance (Q) and the side frictional
accumulated rotation (Rosquoët et al. 2007). Hence, there will resistance (F) (Briaud et al. 1983; Smith 1987). The lateral dis-
always be a net accumulated rotation for balanced two-way loading. placement (d ) of the pile will cause a very complex strain field in
Further, Fig. 4(a) illustrates that the rate of accumulated rotation the soil around it [e.g., refer to Smith (1987) and Zhang et al.
(i.e., a1 value) is much greater for unbalanced two-way loading (2005)]. Fig. 5 shows schematically the idealized stress field
( z c = –0.5), when compared with one-way, partial one-way, and around the pile for a particular depth (z) due to the soil–pile
balanced two-way loading ( z c = 0, 0.5, and –1). This observation interaction.
supports the opinions of LeBlanc et al. (2010) and Zhu et al. (2013) Several theoretical methods to predict the behavior of short rigid
but generally contradicts the trends documented by Long and piles subjected to lateral loading in cohesionless soils have been
Vanneste (1994), Lin and Liao (1999), and Peng et al. (2011). In presented (Brinch Hansen 1961; Broms 1964; Petrasovits and
terms of the accumulated displacement of the monopile, Fig. 4(b) Award 1972; Meyerhof et al. 1981). These methods assume some
tentatively indicates a linear relationship (approximately passing form of LSR distribution profile along the pile embedment length,

© ASCE 04016040-6 Int. J. Geomech.

Int. J. Geomech., 04016040


Downloaded from ascelibrary.org by University of Engineering and Technology, Lahore on 05/03/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 5. Schematic diagram of soil–pile interaction: (a) long section showing rigid pile under lateral loading; (b) cross section showing initial and dis-
placed positions of pile and induced stress regime for depth z

Fig. 6. Different options suggested/adopted for LSR distribution profile along embedded length of rigid pile: (a) Brinch Hansen (1961); (b) Broms
(1964); (c) Petrasovits and Award (1972); (d) Meyerhof et al. (1981); (e) Adams and Radhakrishna (1973); (f) Meyerhof and Sastry (1985); (g) Joo
(1985); (h) Prasad and Chari (1999); (i) Chari and Meyerhof (1983)

which is considered to be uniform across the pile width (Prasad and along the embedded length of the model pile were determined
Chari 1999). Different profile shapes for the LSR distribution by using Eqs. (6) and (7).
along the embedded length of rigid piles have been suggested/
adopted by different researchers, as shown in Fig. 6, and these EIɛ
M ðzÞ ¼ (6)
are considered in the present investigation. c
The structural behavior of a rigid pile under lateral loading is
defined by its rotation as a rigid body, rather than bending (flex- d2 ðMÞ
ing) in the case of long flexible piles. However, because the pðzÞ ¼ (7)
dz2
stiffness of the model pile was finite, some bending strains were
measured at the microstrain level by the strain gauges (bonded
where c (= D/2) = outer radius of the pile; EI = bending stiff-
to its outer wall surface) located at discrete points along its em-
ness of the pile; and « = strain value measured for a particular
bedded length. In the present experimental setup, this was facili-
tated using a System-7000 (Vishay Precision Group, Malvern, depth (z) along the pile embedment length.
PA) data-acquisition system. Bending strain data are usually
used for evaluations/interpretations related to the pile’s bending
moment profile and the LSR and deformations occurring in the Variation in LSR under Static Lateral Loading
soil over the pile embedment length (Arshad and O’Kelly
2014). In the present investigation, the bending moment MðzÞ Fig. 7 shows the shapes of the derived LSR profiles, including the
distribution profile and the LSR pðzÞ (force/length) variation values of LSR (force/length) mobilized along the pile embedment

© ASCE 04016040-7 Int. J. Geomech.

Int. J. Geomech., 04016040


Downloaded from ascelibrary.org by University of Engineering and Technology, Lahore on 05/03/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. Variation of LSR along embeded length of monopile for applied lateral loads at the pilehead of (a) 40 N, (b) 100 N, and (c) 140 N

© ASCE 04016040-8 Int. J. Geomech.

Int. J. Geomech., 04016040


length for three different levels of static (monotonic) applied lateral patterns suggested by Adams and Radhakrishna (1973), Chari and
load [40, 100, and 140 N shown in Figs. 7(a–c), respectively]. From Meyerhof (1983), Joo (1985), Meyerhof and Sastry (1985), and
these figures, it can be concluded that the shapes of the generated Prasad and Chari (1999). These profiles propose that a reversal of
LSR profiles, and the LSR values themselves, are markedly depend- LSR occurs at some distance above the toe of the pile (Fig. 6), i.e.,
ent on the degree of the polynomial functions employed. For at its point of rotation. It should be noted that the accuracy of model-
instance, 3rd-order (piecewise) and 5th-order polynomial functions ing the LSR profiles by using the bending strain data depends on
of the bending strain data generated LSR profiles roughly similar to both the quality and quantity of the strain data.
the one proposed by Broms (1964), i.e., increasing linearly with The lateral displacement of the monopile from its initial vertical
depth, only on one side [see Fig. 6(b)]. Fourth-order polynomial alignment was measured by using two horizontally mounted dis-
placement transducers that contacted against the shaft of the pile
Downloaded from ascelibrary.org by University of Engineering and Technology, Lahore on 05/03/16. Copyright ASCE. For personal use only; all rights reserved.

functions of the bending strain data followed the LSR profile pat-
terns suggested by Adams and Radhakrishna (1973), Chari and that protruded above the finished sand bed surface in the steel tank
Meyerhof (1983), Joo (1985), Meyerhof and Sastry (1985), and (Figs. 1 and 2). From these readings and knowing the vertical sepa-
Prasad and Chari (1999), i.e., curvilinear, bending back on them- ration of the transducers, the depth to the point of rotation of the pile
selves [see Fig. 6(e–i)]. (measured from the sand bed surface level) was determined through
the use of the trigonometric relationship between similar triangles.
Sample results are given in Fig. 9, which illustrates that the point of
Variation in LSR and Point of Rotation of Monopile rotation of the model pile fluctuated between approximately 275
under Repeated Lateral Loading and 325 mm, along its embedded length of 360 mm. This range of
fluctuation for the depth to the point of rotation of the pile approxi-
The bending strains were also measured along the pile embedment mately matches with the point of reversal of the LSR profiles for the
length during the series of repeated lateral loading tests. For different loading schemes investigated (see Fig. 8).
N < 6,000, the bending strains either remained constant at very small
values or showed only very small increases with the number of load Possible Limitations of This Investigation
cycles (their values always remained below 100 m m). This experi-
mental evidence confirms the rigid-body rotation response of the The experimental model developed in the present investigation is
model pile [i.e., not bending (flexing) associated with long flexible based on experimental data obtained for a narrow initial density
piles]. index range of 70–74% for the sand beds and also considered only
The bending moment and the values of LSR were calculated one particular pile diameter and embedment depth. However, the
using Eqs. (6) and (7). Sample results obtained for LSR profiles basic framework of the model is not expected to change for other
deduced on the basis of 4th-order polynomial functions of the bend- density index values of the sand beds. Ideally, 1g testing should be
ing strain data are presented in Fig. 8. Fourth-order polynomial performed using sand beds prepared in a (very) loose state, although
functions were selected (over 3rd- and 5th-order polynomials) for it is not uncommon to find comparable experimental studies
this discussion because these LSR profiles were found to match reported for similar or higher dense states (e.g., Peng et al. 2011;
closely with actual LSR profiles measured by using pressure trans- Cuellar et al. 2012; Nasr 2014; Arshad and O’Kelly 2016c), prob-
ducers (Adams and Radhakrishna 1973; Chari and Meyerhof 1983; ably to explore the generic response of the monopile under load-
Joo 1985; Meyerhof and Sastry 1985; Prasad and Chari 1999). On term repeated loading. Further, because homogeneous saturation of
comparing the sample results with the options presented in Fig. 6, it the sand beds would have been challenging, the entire experimental
can be concluded that the experimental results based on 4th-order program investigated the performance of the model pile installed in
polynomial functions of the bending strain data follow the profile dry sand beds.

Fig. 8. LSR distribution profiles for various repeated lateral loading schemes

© ASCE 04016040-9 Int. J. Geomech.

Int. J. Geomech., 04016040


Downloaded from ascelibrary.org by University of Engineering and Technology, Lahore on 05/03/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 9. Variation in depth to point of rotation of model pile for repeated lateral loading tests

Another limitation of this investigation might be that the The test results also show that the cyclic stiffness of the soil−pile sys-
number of load cycles applied in each model test was limited to tem always increased under repeated lateral loading, probably as a
6,000, whereas for field conditions, an OWT monopile might be result of densification of the sand around the pile, whereas the abso-
subjected to as much as N = 107 for the fatigue limit state lute secant stiffness always had a decreasing trend due to the continu-
(Germanischer Lloyd Industrial Services 2005). The application ous accumulating rotation of the pile. The proposed logarithmic law,
of such a large number of load cycles in the present investigation which involved an additive approach, captured the experimental
would be very time consuming, even for a single test. However, accumulated rotation against number of load cycles relationship
from the experimental data gathered, the authors observed that more precisely, when compared with the power law approach. In the
the trend in the rotation behavior of the model pile under development of the experimental model, only a narrow range of den-
repeated lateral loading had essentially become stable by N = sity index values for the sand beds was considered. Further investiga-
6,000; i.e., its rotation might continue to increase with increas- tions of this framework for other values of sand density through ex-
ing N, but at a diminishing rate. Any quick change in the curve perimental and numerical studies are recommended. As a general
describing the rotation against number of load cycles relation- observation, interpretations of the distribution profiles and values of
ship, as had occurred during the application of the first 500 load the LSR based on curve-fitting of measured bending strain data can
cycles, seems unlikely. highly misleading/unreliable.

Conclusions Acknowledgments

This paper has identified various limitations and discrepancies asso- The first author gratefully acknowledges a postgraduate research
ciated with the design and analysis methodologies currently used award received from Trinity College Dublin. The authors thank
for large-diameter monopiles subjected to repeated lateral loading. the reviewers for many helpful comments.
A series of tests was performed on a model monopile to understand
its response under long-term repeated lateral loading. At z b ¼ 0:42 Notation
(serviceability limit state), unbalanced two-way loading ( z c ¼
0:5) was found to generate approximately 125–155% more rota- The following symbols are used in this paper:
tion of the monopile (normalized values) compared with that pro- c ¼ outer radius of pile;
duced for one-way loading ð z c ¼ 0Þ for N = 500–6,000. The corre- D ¼ outer diameter of pile;
sponding increase for z b ¼ 0:28 (fatigue limit state) was 350–450%. d50 ¼ mean effective grain size;

© ASCE 04016040-10 Int. J. Geomech.

Int. J. Geomech., 04016040


EI ¼ bending stiffness; supported offshore wind turbines.” Int. J. Phys. Modell. Geotech.
F ¼ side frictional resistance mobilized along pile 11(2), 58–68.
embedment length; Briaud, J. L., Smith, T. D, and Meyer, B. J. (1983). “Using the pressure-
H ¼ lateral load; meter curve to design laterally loaded piles.” Proc., 15th Annual
L ¼ pile embedment length; Offshore Technology Conference, Houston, TX. Vol. 1, Society of
Petroleum Engineers, Houston, 495–502.
LSR ¼ lateral soil resistance;
Brinch Hansen, J. (1961). The ultimate resistance of rigid piles against
Pmax ¼ maximum lateral load at pilehead during repeated transversal forces, Bulletin No. 12, Danish Geotechnical Institute,
lateral loading; Copenhagen, Denmark, 5–9.
Pmin ¼ minimum lateral load at pilehead during repeated Broms, B. B. (1964). “Lateral resistance of piles in cohesionless soils.” J.
lateral loading;
Downloaded from ascelibrary.org by University of Engineering and Technology, Lahore on 05/03/16. Copyright ASCE. For personal use only; all rights reserved.

Soil Mech. Found. Eng., 90(SM3), 123–156.


Pu ¼ ultimate static lateral load-carrying capacity of Byrne, B. W., and Houlsby, G. T. (2003). “Foundations for offshore
pile; wind turbines.” Philos. Trans. R. Soc. London, Ser. A, 361(1813),
pðzÞ ¼ LSR profile (force/length); 2909–2930.
MðzÞ ¼ bending moment distribution profile; Chari, T. R., and Meyerhof, G. G. (1983). “Ultimate capacity of rigid sin-
N ¼ number of load cycles; gle piles under inclined loads in sand.” Can. Geotech. J. 20(4),
OWT ¼ offshore wind turbine; 849–854.
Q ¼ frontal normal resistance mobilized along pile Cuellar, P. (2011). “Pile foundations for offshore wind turbines: Numerical
and experimental investigations on the behaviour under short-term and
embedment length;
long-term cyclic loading.” Ph.D. thesis, Technical Univ. of Berlin,
R2 ¼ coefficient of determination;
Berlin.
z ¼ particular depth below sand bed surface level; Cuellar, P., Georgi, S., Bæßler, M., and Rucker, W. (2012). “On the
a1 ¼ model parameter for logarithmic law[(= fn ( z c )]; quasi-static granular convective flow and sand densification around
b 1 ¼ model parameter for logarithmic law [= fn ( z b )]; pile foundations under cyclic lateral loading.” Granular Matter,
a2 ; b 2 ¼ models for power law; 14(1), 11–25.
d ¼ lateral displacement; Davie, J. R., and Sutherland, H. B. (1978). “Modeling of clay uplift resist-
« ¼ strain value corresponding to depth z along embed- ance.” J. Geotech. Eng., 104(6), 755–760.
ded length of pile; DNV (Det Norske Veritas) (2011). DNV–OS–J101: Design of offshore
u N ¼ accumulated rotation of monopile for N load wind turbine structures. DNV, Oslo, Norway.
cycles; Dong, P., Newson, T. A., Davies, M. C. R., and Davies, P. A. (2001).
u 0 ¼ rotation of monopile for end of first load cycle; “Scaling laws for centrifuge modelling of soil transport by turbulent
Du ðNÞ ¼ rotation produced from second load cycle up to fluid flows.” Int. J. Phys. Modell. Geotech. 1(1), 41–45.
El Sawwaf, M. (2006). “Lateral resistance of single pile located near
Nth load cycle;
geosynthetic reinforced slope.” Geotech. Geoenviron. Eng.10.1061/
z b ¼ parameter to define cyclic character of repeating (ASCE)1090-0241(2006)132:10(1336), 1336–1345.
loads; Franke, E., and Muth, G. (1985). “Scale effect in 1-g model tests on
z c ¼ parameter to define cyclic character of repeating horizontally loaded piles.” Proc., 11th Int. Conf. on Soil Mechanics
loads; and Foundation Engineering, Vol. 2, A. A. Balkema, Rotterdam,
1g ¼ normal gravitational condition; Netherlands, 1011–1014.
1w ¼ one-way (loading); and Germanischer Lloyd Industrial Services. (2005) Guidelines for the certifica-
2w ¼ two-way (loading). tion of offshore wind turbines, Germanischer Lloyd WindEnergie
GmbH, Hamburg, Germany.
Haiderali, A. E., and Madabhushi, G. S. P. (2013). “Evaluation of the p-y
References method in the design of monopiles for offshore wind turbines.” Proc.,
Offshore Technology Conf., Vol. 3, Society of Petroleum Engineers,
Adams, J. I., and Radhakrishna, H. S. (1973). “The lateral capacity of deep Houston, 1824–1844.
augured footings.” Proc., 8th Int. Conf. on Soil Mechanics and Haiderali, A. E., Lau, B. H., Haigh, S. K., and Madabhushi, G. S. P. (2014).
Foundation Engineering, Vol. 2, Strojizdat, Moscow, 1–8. “Lateral response of monopiles using centrifuge testing and finite ele-
API (American Petroleum Institute) (2010). API RPA2: Recommended ment analysis.” Proc., 8th Int. Conf. on Physical Modelling in
practice for planning, designing and constructing fixed offshore plat- Geotechnics, Vol. 2, C. Gaudin and D. J. White, eds., CRC Press, Boca
forms—Working stress design, 22nd ed., Washington, DC. Raton, FL, 743–749.
Arshad, M., and O'Kelly, B. C. (2013). “Offshore wind-turbine structures: a Hettler, A. (1981). “Verschiebungen starrer und elastischer gründungskörper
review.” Proc. Inst. Civ. Eng. Energy, 166(4), 139–152. in sand bei monotone und zyklischer belastung.” Ph.D. thesis, Karlsruhe
Arshad, M., and O’Kelly, B. C. (2014). “Development of a rig to study model Institute of Technology, Karlsruhe, Germany (in German).
pile behaviour under repeating lateral loads.” Int. J. Phys. Modell. Joo, J. S. (1985). “Behavior of large scale rigid model piles under inclined
Geotech. 14(3), 54–67. loads in sand.” M.S. engineering thesis, Memorial Univ. of
Arshad, M., and O'Kelly, B. C. (2016a). “Piled-cruciform attachment to Newfoundland, St. John’s, Newfoundland, Canada.
monopile head reduces deflection.” Proc., Inst. Civ. Eng. Geotech. Eng., Kelly, R. B., Houlsby, G. T., and Byrne, B. W. (2006). “A comparison of
169, in press. field and laboratory tests of caisson foundations in sand and clay.”
Arshad, M., and O’Kelly, B. C. (2016b). “Analysis and design of monopile Geotechnique, 56(9), 617–626.
foundations for offshore wind-turbine structures.” Marine Georesour. Klinkvort, R. T., and Hededal, O. (2013). “Lateral response of monopile
Geotechnol., 34, in press. supporting an offshore wind turbine.” Proc. Inst. Civil Eng. Geotech.
Arshad, M., and O'Kelly, B. C. (2016c). “Reducing monopile rotation under Eng., 166(2), 147–158.
lateral loading in sandy soils.” Geomech. Geoeng., 11, in press. Lai, S. (1989). “Similitude for shaking table test on soil-structure-fluid
Bhattacharya, S., and Adhikari, S. (2011). “Experimental validation of soil- model in 1-g gravitational field.” Soils Found. 29(1), 105–118.
structure interaction of offshore wind turbines.” Soil Dyn. Earthquake Lau, B. H., Lam, S. Y., Haigh, S. K., and Madabhushi, G. S. P. (2014).
Eng. 31(5–6), 805–816. “Centrifuge testing of monopile in clay under monotonic loads.” Proc.,
Bhattacharya, S., Lombardi, D., and Muir Wood, D. (2011). 8th Int. Conf. on Physical Modelling in Geotechnics, Vol. 2, C. Gaudin
“Similitude relationships for physical modelling of monopile- and D. J. White, eds., CRC Press, Boca Raton, FL, 689–695.

© ASCE 04016040-11 Int. J. Geomech.

Int. J. Geomech., 04016040


LeBlanc, C. (2009). “Design of offshore wind turbine support structures: Peng, J., Clarke, B., and Rouainia, M. (2011). “Increasing the resistance of
Selected topics in the field of geotechnical engineering.” Ph.D. thesis, piles subject to cyclic lateral loading.” Geotech. Geoenviron. Eng., 10
Aalborg Univ., Aalborg, Denmark. .1061/(ASCE)GT.1943-5606.0000504, 977–982.
LeBlanc, C., Houlsby, G. T., and Byrne, B. W. (2010). “Response of stiff piles Petrasovits, G., and Award, A. (1972). “Ultimate lateral resistance of a rigid
in sand to long-term cyclic lateral loading.” Geotechnique, 60(2), 79–90. pile in cohesionless soil.” Proc., 5th European Conf. on Soil Mechanics
Lin, S-S., and Liao, J-C. (1999). “Permanent strains of piles in sand due to and Foundation Engineering, Vol. 3, Spanish Society of Soil
cyclic lateral loads.” Geotech. Geoenviron. Eng.10.1061/(ASCE)1090 Mechanics and Foundations, Madrid, 407–412.
-0241(1999)125:9(798), 798–802. Prasad, Y. V. S. N., and Chari, T. R. (1999). “Lateral capacity of model rigid
Little, R. L., and Briaud, J. L. (1988). “Full scale cyclic lateral load tests on piles in cohesionless soils.” Soils Found., 39(2), 21–29.
six single piles in sand.” Rep. GL–88–27, Geotechnical Division, Civil Rao, S. N., Ramakrishna, V. G. S. T., and Raju, G. B. (1996).
Engineering Dept., Texas A&M Univ., College Station, TX. “Behavior of pile-supported dolphins in marine clay under lateral
Downloaded from ascelibrary.org by University of Engineering and Technology, Lahore on 05/03/16. Copyright ASCE. For personal use only; all rights reserved.

Long, J. H., and Vanneste, G. (1994). “Effects of cyclic lateral loads on piles loading”. J. Geotech. Eng., 10.1061/(ASCE)0733-9410(1996)122:
in sand.” J. Geotech. Eng.10.1061/(ASCE)0733-9410(1994)120:1(225), 8(607), 607–612.
225–244. Rosquoët, F., Thorel, L., Garnier, J., and Canepa, Y. (2007). “Lateral cyclic
Meyerhof, G. G., Mathur, S. K., and Valsangkar, A. J. (1981). “Lateral re- loading of sand-installed piles.” Soils Found., 47(5), 821–832.
sistance and deflection of rigid walls and piles in layered soils.” Can. Sedran, G., Stolle, D. F. E., and Horvath, R. G. (2001). “An investigation of
Geotech. J., 18(2), 159–170. scaling and dimensional analysis of axially loaded piles.” Can. Geotech.
Meyerhof, G. G., and Sastry, V. V. R. N. (1985). “Bearing capacity of rigid J., 38(3), 530–541.
piles under eccentric and inclined loads.” Can. Geotech. J., 22(3), Smith, T. D. (1987). “Pile horizontal modulus values.” J. Geotech. Eng., 10
267–276. .1061/(ASCE)0733-9410(1987)113:9(1040), 1040–1044.
Muir Wood, D., Crewe, A. J., and Taylor, C. A. (2002). “Shaking table Tomlinson, M. J. (2001). Foundation design and construction, 7th ed.,
testing of geotechnical models.” Int. J. Phys. Modell. Geotech., 2(1), Pearson, Essex, U.K.
1–13. Verdure, L., Garnier, J., and Levacher, D. (2003). “Lateral cyclic loading of
Nasr, A. M. A. (2014). “Experimental and theoretical studies of laterally single piles in sand.” Int. J. Phys. Modell. Geotech., 3(3), 17–28.
loaded finned piles in sand.” Can. Geotech. J., 51(4), 381–393. Wichtmann, T., Rondón, H. A., Niemunis, A., Triantafyllidis, Th., and
Niemunis, A., Wichtmann, T., and Triantafyllidis, T. (2005). “A high-cycle Lizcano, A. (2010). “Prediction of permanent deformations in pave-
accumulation model for sand.” Comput. Geotech. 32(4), 245–263. ments using a high-cycle accumulation model.” Geotech. Geoenviron.
O’Kelly, B. C., and Arshad, M. (2016) “Chapter 20: Offshore wind turbine Eng., 10.1061/(ASCE)GT.1943-5606.0000275, 728–740.
foundations—Analysis and design.” Offshore wind farms: technologies, Zhang, L., Silva, F., and Grismala, R. (2005). “Ultimate lateral resistance to
design and operation, C. Ng and L. Ran, eds., Woodhead Publishing, piles in cohesionless soils.” Geotech. Geoenviron. Eng., 10.1061/
Cambridge, U.K., 589–610. (ASCE)1090-0241(2005)131:1(78), 78–83.
Ovesen, N. K. (1979). “Panel discussion in session 9: The use of physical mod- Zhu, B., Byrne, B. W., and Houlsby, G. T. (2013). “Long-term lateral cyclic
els in design.” Proc., 7th European Conf. on Soil Mechanics and Foundation response of suction caisson foundations in sand.” Geotech. Geoenviron.
Engineering, Vol. 4, British Geotechnical Society, London, 319–323. Eng., 10.1061/(ASCE)GT.1943-5606.0000738, 73–83.

© ASCE 04016040-12 Int. J. Geomech.

Int. J. Geomech., 04016040


Geotechnical Engineering Proceedings of the Institution of Civil Engineers

http://dx.doi.org/10.1680/jgeen.15.00001
Piled-cruciform attachment to monopile Paper 1500001
Received 13/02/2015 Accepted 09/02/2016
head reduces deflection
Keywords: foundations/models (physical)/offshore engineering
Arshad and O’Kelly

ICE Publishing: All rights reserved

Piled-cruciform attachment to
monopile head reduces
deflection
&
1 Muhammad Arshad BSc, MSc, ME, PhD &
2 Brendan C. O’Kelly PhD, FTCD, CEng, CEnv, MICE
Assistant Professor, Department of Geological Engineering, Associate Professor, Department of Civil, Structural and Environmental
University of Engineering and Technology, Lahore, Pakistan; Engineering, Trinity College Dublin, Dublin, Ireland
formerly PhD candidate, Department of Civil, Structural and
Environmental Engineering, Trinity College Dublin, Dublin, Ireland
(corresponding author: arshadm@tcd.ie)

1 2

Much critical infrastructure, including bridges, wind turbine structures, dolphins and some other ocean engineering
structures, is supported on large-diameter rigid monopiles. For such structures, compared with the gravitational loads,
cyclic lateral loading may often be more critical for the analysis and design. The lateral load-carrying capacity of a pile
depends on its geometry (dimensions), the soil properties and type of loading. In order to increase its lateral load-
carrying capacity, it is necessary either to change the properties of the near-surface layers of soil or to change its
geometry. This paper presents model studies investigating a novel technique to limit the lateral deflection (rotation)
of a monopile under long-term cyclic lateral loading. The technique provides enhanced restraint of the monopile
through the installation of four shorter piles, arranged in a cruciform, which attach to the head of the central
monopile by way of a grillage. Different aspects of this modification, including its fabrication and attachment to the
monopile, are presented. Its efficiency in reducing the monopile rotation under cyclic lateral loading is evaluated
through a comprehensive testing programme, with reasonably encouraging results.

Notation conditions, water depth, anticipated loading and financial con-


a monopile rotation straints (Malhotra, 2010). Large-diameter monopiles are a fre-
b rotation of monopile with piled-cruciform attach- quently used foundation system for offshore wind turbines and
ment in situ other offshore structures. Offshore wind farms often contain
D outer diameter of pile can comprise many hundreds of turbines supported at heights
d10 effective grain size for 10% passing by mass of typically 30–80 m above mean sea level. The preferred foun-
d30 effective grain size for 30% passing by mass dation type for these tall structures in water depths of up to
d50 mean effective grain size 30 m is large-diameter monopiles, owing to their ease of con-
d60 effective grain size for 60% passing by mass struction and installation. These monopiles are subjected to
EI bending stiffness large cyclic lateral and moment loads, in addition to axial
kh initial coefficient of subgrade reaction loads, as documented by LeBlanc et al. (2010), Bhattacharya
L monopile embedment length et al. (2011) and Cuéllar (2011). However, their lateral load-
LSDP embedment length of small diameter pile carrying resistance may not be sufficient to withstand pro-
Pu ultimate static lateral load-carrying capacity longed impact under large wind and wave loading. A solution
ηL dimensionless embedment length of monopile may be achieved by simply increasing the physical dimensions
of the monopile, but this may not be economically or practi-
1. Introduction cally feasible. Thus, a range of different foundation solutions
The preferred foundation solution for a particular offshore in place of, or modifications of, the monopile are being investi-
structure depends, among other factors, on the local soil gated by different researchers.

1
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

Modification of the conventional monopile by the attachment (Achmus et al., 2009). Monopile foundations are generally
of ‘wings’ close to the pilehead, in order to increase the lateral used in shallow water depths (i.e. typically < 30 m), generally
load-carrying capacity and stiffness of the foundation system becoming too flexible for water depths of between  30 and
for weaker soil conditions (close to the mudline), has been 40 m, in which case monopiles fitted with guy wires or tripod
investigated by many researchers using small-scale tests in sand solutions are considered as economical alternatives. For greater
under the normal gravitational (1g) condition (Dührkop and water depths (> 40 m), time-consuming installation and the
Grabe, 2008; Nasr, 2014; Peng et al., 2011) and in centrifuge effects of soil degradation (‘potholing’) that occurs in-service
facilities (Bienen et al., 2012). These researchers found that, at mudline level around the pile make monopile foundations
with the wings attached, the pilehead deflection substantially prohibitive (Irvine et al., 2003). Other foundation options, as
reduced (by  50−70%) and the ultimate lateral load-carrying discussed by Arshad and O’Kelly (2013, 2016a) and O’Kelly
capacity increased by up to 80%, depending on the length of and Arshad (2016), are then considered to be viable. The servi-
the wings compared with the length of the monopile, the shape ceability limit state is largely determined by the lateral deflec-
of the wings and the soil properties. Another alteration, com- tion (rotation) response of the monopile under many millions
prising a monopile combined with a footing base, has also been of load cycles; for example, over the service life of a 2 MW
proposed. Initial model tests performed in sand at 1g were OWT structure, 107 lateral load cycles of 1·4 MN magnitude
reported by Stone et al. (2007), with apparently promising (corresponding to the fatigue loading for design) are expected
results, suggesting that the additional rotation restraint provided to occur (Germanischer Lloyd, 2005).
by the footing can result in a stiffer lateral response and greater
ultimate lateral load-carrying capacity. Arshi et al. (2013) per- The monopile must mobilise sufficient soil resistance over its
formed tests in sand at 1g by adding skirts of different lengths embedded length to transfer all types of applied loads to the
(depths) to these piled footings. Their results indicated that surrounding soil, with adequate safety factors, and prevent toe
increasing the skirt length tends to increase the ultimate lateral ‘kick’ (displacement of the pile base) and excessive deflection/
load-carrying capacity of the foundation system by about 50% rotation of the pile itself. According to current practice, mono-
compared with the non-skirted hybrid system. More recently, piles are analysed for the axial loads only to determine their
Arshad and O’Kelly (2016b) reported 1g tests in sand that bearing capacity and settlement responses, and then for the
investigated the use of concentric rings of small-diameter piles lateral loads only to determine their lateral load-carrying
(SDPs) installed centrally around the model monopile. Their capacity and flexural behaviour (Karthigeyan et al., 2006;
results showed that the rotation of the monopile, investigated Moayed et al., 2012). Compared with the axial loads, the
under a range of cyclic lateral loading scenarios, was reduced lateral loads are considered to be governing, as mentioned in
by 40–65% owing to the enhanced confinement and densifica- several design guidelines (API, 2010; DNV, 2011;
tion of the sand test bed provided by the presence of the SDPs. Germanischer Lloyd, 2005) and documented by many
researchers (Achmus, 2010; Bhattacharya et al., 2013; Carswell
This paper presents an experimental investigation performed et al., 2015; Haiderali et al., 2013; Kuo et al., 2012; Leblanc
to explore the possibility of pilehead modification to reduce et al., 2010; Lombardi et al., 2013; Malhotra, 2010; Nicolai
the accumulated rotation of a monopile under long-term cyclic and Ibsen, 2014; Peng et al., 2011; Zhu et al., 2013). In other
lateral loading. The novel solution proposed comprises four words, the required diameter, wall thickness and embedment
shorter SDPs, arranged in a cruciform, which attach by way of length of the monopile is generally dictated by the applied
a grillage to the pilehead at the mudline level. From a review lateral loads and moments. Hence, the experimental work pre-
of the literature, this would appear to be the first study of its sented in this paper focuses on the enhancement of the lateral
kind in relation to the proposed set-up for deep foundation stability of the conventional monopile, with a novel modifi-
structures. On the basis of encouraging results, it can be cation to the pilehead proposed, namely four SDPs arranged
expected that this novel technique may prove to be a viable sol- in a cruciform that attach by way of a grillage to the pilehead.
ution to enhance the serviceable life of structures supported by
monopiles that are subjected to long-term cyclic lateral loading, 2.2 Geometric details of proposed arrangement
such as offshore wind-turbine structures (OWTs). Figure 1 shows a schematic diagram of the proposed arrange-
ment at reduced scale, which consists of a central ‘split-able’
ring with four radial steel arms, each fitted with a smaller
2. Development of the proposed
diameter ring (sleeve ring) at its far end. The two halves of the
arrangement
central ring, manufactured from 30 mm wide  2 mm thick
2.1 Governing loading for OWT monopile foundation steel strip, are secured together by way of their collars, using
system M4 nut–bolts, clamping around the head of the monopile
Monopile foundations for offshore wind-turbine structures are (D = 53 mm). Four radial struts welded to the central ring,
typically manufactured from steel tubular sections with an each having length, width and thickness dimensions of 85, 20
outer diameter (D) of up to 7·5 m, wall thicknesses of up to and 2 mm, respectively, were arranged in a cruciform, making
150 mm and embedment depths of between 15 and 30 m the four arms. The far end of each of these arms was

2
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

125 mm

125 mm 125 mm

19 mm 53 mm
Central ring to house 53 mm
diameter monopile
M4 nut–bolt passing through
A the side collar A

20 mm wide and 2 mm thick


mild steel arm
(a)
M4 nut–bolt
Smaller diameter ring to
house 19 mm diameter pile

Jacking screw

30 mm 20 mm

(b)

Figure 1. Schematic diagram of cruciform attachment for


monopile: (a) plan view; (b) side view (section A–A)

connected by an M4 nut–bolt to a sleeve ring, manufactured et al., 2011; Cuéllar et al., 2012; LeBlanc et al., 2010;
from 25 mm wide  2 mm thick steel strip. The connections Rosquoët et al., 2007). For sandy soil, it can be argued that
between the radial arms and sleeve rings allowed changes to be the zone of significant influence remains limited to within
made to the inclination of the SDPs that were housed in these 2–3D measured from the monopile axis, as evident from the
rings. The sleeve rings were equipped with jacking screws (see formation of a cone of depression observed at the sand bed
Figure 1) that allowed the required adjustments to be made to surface level around the monopile during 1g testing (Brown
the solid SDPs (D = 19 mm; i.e. 36% of the monopile diam- et al., 1988; Cuéllar et al., 2012). For lateral loading, the upper
eter) during their installation in the sand beds (described part of the soil deposit around the monopile is more critical
later). than the lower part, owing to the greater level of deflection oc-
curring closer to the pilehead (Nasr, 2014; Zhang et al., 2005).
Figure 2 shows the arrangement centred on the monopile, with
the four solid SDPs aligned vertically and attached by way of Conceptually, the piled-cruciform attachment is designed to
the grillage to the pilehead. Hereafter, the central split-able transfer some of the applied lateral load away from the zone
ring (holding the monopile) and the four radial arms holding of significant influence for the monopile. Under the action of
the SDPs are collectively termed as the piled-cruciform the applied lateral loading/moment, the monopile and piled-
arrangement. cruciform attachment tend to deflect (rotate with respect to
their initial vertical alignment), with resistance provided by the
2.3 Working mechanism of piled-cruciform passive pressures (forces) mobilised along the embedded
arrangement lengths of the monopile and four SDPs (the latter are trans-
The soil around the monopile is influenced by the cyclic lateral mitted by way of the cruciform arms to the pilehead). Through
loading (Bhattacharya and Adhikari, 2011; Bhattacharya this arrangement, the stress intensity is reduced in the

3
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

the resulting axial and lateral deflections and rotation response


of the pile (Figure 4) are presented in the paper by Arshad and
O’Kelly (2014).
Monopile
The model pile was manufactured from smooth brass tubing,
Central split-able ring
having an overall length of 540 mm, outer diameter of
Arm 53·0 mm and wall thickness of 0·8 mm, which produced a
Smaller bending stiffness (EI) value of 4·33 kN m2. Its lower end was
ring closed using a 3 mm thick brass plate in order to represent a
fully plugged tubular pile. Geometrically, the pile set-up, with
an embedded length (L) to outer diameter (D) ratio of 6·8 at
the start of each loading test performed, is categorised as a
short ‘rigid’ pile, which encompasses L/D ratios of up to 10
SDP (DNV, 2011; Peng et al., 2011; Tomlinson, 2001). To verify
this, the pile’s rigidity was evaluated based on its value of the
dimensionless embedment length, ηL (Broms, 1964). The coef-
ficient η is calculated as

 1=5
kh
1: η¼
EI

where kh is the initial coefficient of subgrade reaction and EI is


the bending stiffness of the monopile.

According to Terzaghi (1955), the value of kh is 19·8 MN/m3


for dense sand (which was the soil investigated in the present
study). A pile is considered to be a short rigid pile for ηL < 2
and a long elastic pile for ηL > 4 (Broms, 1964; Chari and
Figure 2. Piled-cruicform arrangement attached to monopile
Meyerhof, 1983). The estimated ηL value of 1·94 for the mono-
pile set-up in the present investigation indicates that the model
pile satisfied the criterion for short rigid piles.
immediate vicinity of the monopile, thereby lowering the strain
developed in the surrounding soil, as compared with the mono- The model pile was partially embedded in dense sand beds pre-
pile alone for the same magnitude of lateral load. Further, pared in a 0·95 m dia. by 0·6 m deep steel tank. These dimen-
instead of having a ‘free’ pilehead, the rotation is restrained sions were chosen to ensure that for the static and cyclic lateral
by the cruciform arm attachment. The proposed set-up also pro- loading scenarios investigated in the present study, the failure
duces a significant increase in the bending stiffness, as compared wedge for the pile would not extend to the tank boundaries.
with the conventional monopile. For instance, considering the With a ratio of tank diameter to pile diameter of almost 18,
structural elements in the present experimental set-up, the side wall boundary effects were not considered to be significant
second moment of area about the centroidal axis of the 53 mm (Davie and Sutherland, 1978; Rao et al., 1996). Similarly, a
diameter hollow monopile was 4·5  104 mm4, with its value soil cushion (sand in the present case) having a thickness of 3–
increased to  8·9  106 mm4 at the mudline level when fitted 4D located below the pile base was considered sufficient to
with the piled-cruciform attachment; that is, 200 times greater. absorb the small vertical stress field (LeBlanc et al., 2010).

3. Testing facility, instrumentation and 4. Sand characterisation and sand bed


model pile preparation, including monopile and SDP
For the present study, the Trinity College Dublin electro- installation
mechanical loading system for testing model piles at 1g All of the tests performed used commercially available air-
(Figure 3) was used. This system is capable of applying many dried Glenview sand, comprising sub-angular to angular
thousands of lateral load cycles on scaled models, with full grains ranging from 0·1 to 1·0 mm in size, with d10, d30, d50
control provided over the loading direction, amplitude, fre- and d60 values of 0·16, 0·22, 0·27 and 0·31 mm, respectively,
quency and waveform shape. Further details on its working giving a coefficient of uniformity value of 1·94 and coefficient
mechanism and capabilities, including the instrumentation of curvature value of 0·98. Furthermore, the sand had
used to measure the lateral loads applied to the pilehead and minimum and maximum dry density values of 1388 and

4
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

L O R

Sand level in steel tank


7
8 3

10
11 4
12
13 5

Figure 3. Schematic diagram of the experimental rig for model articulated arm; 10, sliding node; 11, left loading hanger, with
pile studies, set up for two-way lateral loading of the pilehead. weight; 12, reference support with two horizontal displacement
Note: 1, drive motor; 2, upper right chain segment; 3, lower right transducers; 13, monopile, shown without piled-cruciform
chain segment; 4, right loading hanger, with weight; 5, steel attachment (adopted from Arshad and O’Kelly (2014))
tank; 6, reaction frame; 7, right load cell; 8, right spring; 9,

Vertical displacement
transducer
Two horizontal displacement
transducers
Pilehead
Right load cell
50 mm
Right spring
Lower right chain segment

90 mm ( = moment arm)
Sand bed surface level

Figure 4. Arrangement of load cells and displacement


transducers at the pilehead (adopted from Arshad and O’Kelly
(2014))

1662 kg/m3, respectively, equating to maximum and minimum The sand was air-pluviated into the steel tank, gently raining
void ratio values of 0·92 and 0·60, respectively. At maximum in six separate layers, each 100 kg in mass. This produced
density, the dry sand had a peak friction angle of 39°, deter- approximately 90 mm thick deposited layers. After the first
mined using 60 mm-square shearbox tests. two layers of sand had been deposited in the tank, the model

5
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

pile was positioned vertically at the centre of the tank, with


temporary support to its head provided by four tensioned hori- Test number Set-up Loading scenario
zontal steel wires that were secured radially from the wall of (refer to Table 2) (test ID)
the tank. Four more layers of sand were then deposited, bring-
1 Reference 2w/30–60/0·4
ing the sand bed to its full depth of 0·54 m (i.e. pile embed-
2 A 2w/30–60/0·4/A
ment length of 0·36 m). During this process, the surface of the
3 B 2w/30–60/0·4/B
deposited sand layer was levelled, as necessary, using a straight
4 C 2w/30–60/0·4/C
edge before depositing the next sand layer. For all of the sand
5 D 2w/30–60/0·4/D
beds prepared in the present investigation, this preparation tech-
6 E 2w/30–60/0·4/E
nique was found to produce sand beds having a dry density of
7 Reference 2w/60–60/0·4
1577 ± 6 kg/m3 (density index range of 70–74%). The resulting
8 A 2w/60–60/0·4/A
‘wished in place’ pile simulated a pre-bored pile; in other
9 B 2w/60–60/0·4/B
words, no displacement would occur in the sand deposit as a
10 C 2w/60–60/0·4/C
result of the pile installation.
11 D 2w/60–60/0·4/D
12 Reference 1w/60/0·25
On completion of the sand bed with pile installed, the split-able
13 A 1w/60/0·25/A
ring attachment was clamped around the pilehead such that its
14 B 1w/60/0·25/B
cruciform arms were located just above the finished sand
15 C 1w/60/0·25/C
surface level in the steel tank. In the next step, four SDPs were
16 Reference 1w/30–60/0·25
inserted through the sleeve rings and pushed into the sand bed
17 A 1w/30–60/0·25/A
to the required embedment depth (Figure 5). Jacking screws on
18 B 1w/30–60/0·25/B
the sleeve rings (see Figure 1) allowed some fine adjustments
before rigidly connecting them to the heads of the SDPs. Table 1. Cyclic lateral loading test programme

5. Testing programme
A comprehensive programme of cyclic lateral loading tests (see two-way loading conditions), with typically 12 000 lateral load
Table 1) was performed on the model pile in the dense sand cycles applied during each test. For one-way loading, the
beds to evaluate the efficiency of the proposed arrangement. applied load ranged between zero and some maximum load
The tests investigated a load amplitude of 60 N, two loading value, whereas for partial one-way loading, the applied load
frequencies of 0·25 and 0·4 Hz, and different loading direc- ranged between some value above zero and the maximum
tions (one-way, partial one-way, balanced and unbalanced load value. With respect to the ultimate static lateral load-
carrying capacity (Pu) of the pile, the load amplitude of 60 N
corresponds to its serviceability limit state ( 42% of Pu
(DNV, 2011)).

1 5 The tests performed are identified as follows: lateral loading


6 direction (1w, one-way; 2w, two-way)/load amplitude (N)/fre-
2 quency (Hz). For instance, 1w/60/0·25 indicates one-way lateral
3 7 loading of the monopile (without the piled-cruciform arrange-
4 8 ment in situ) for an amplitude of 60 N and frequency of
0·25 Hz. 1w/30–60/0·25 indicates one-way lateral loading, with
the load fluctuating from 30 to 60 N from the same side at
0·25 Hz. 2w/30–60/0·4 indicates unbalanced two-way lateral
loading, 30 N in one direction and 60 N in the other, applied
at a frequency of 0·4 Hz.

Table 2 shows the different formations of the piled-cruciform


arrangement investigated (labelled set-ups A–E), namely:
Figure 5. Piled-cruciform arrangement attached to head of (a) ratio of embedment length of the SDPs (LSDP) to that of
monopile installed in dry sand bed. Note: 1, sand bed surface; the monopile (L); (b) inclination of the SDPs from the vertical
2, protruding head of SDP; 3, sleeve ring; 4, central ‘split-able’ direction; (c) orientation of the cruciform grillage with respect
ring; 5, steel tank perimeter; 6, protruding head of central to the line of action of the applied loading. For all of the static
monopile; 7, cruciform arm; 8, miniature load cell attached to pile and cyclic tests performed, the lateral loading was applied to
shaft at a distance of 90 mm above the sand bed surface level the side of the model pile at an elevation of 90 mm above the
sand bed surface level.

6
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

Set-up SDP length (LSDP) and Offset of cruciform arms Comment


inclination to vertical with respect to line of action
direction of applied loading

Cruciform arms are


aligned with the line of
action of the cyclic
lateral loading.
A
Vertical SDPs are half
the length of the central
monopile (L) – that is,
LSDP /L = 0·5.

As per set-up A, apart


from the SDPs being
one-third of the
monopile length – that is,
B LSDP /L = 0·33.

Cruciform arms are


offset by 45° from the
line of action of the
cyclic lateral loading.
C
Vertical SDPs are one-
third of the monopile
length; that is, LSDP /L =
0·33.

As per set-up B, apart


from the SDPs being
inclined at 15° to the
vertical direction.
D

As per set-up C, apart


from the SDPs being
inclined at 15° to the
vertical direction.
E

Table 2. Different set-ups of the piled-cruciform attachment


investigated

7
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

6. Ultimate static lateral load-carrying 2006; Peng et al., 2011; Uncuoğlu and Laman, 2011) occurring
capacity of monopile at the sand bed surface level. Hence, in the present investi-
Static lateral loads were applied in small increments of 10 N to gation, the value of Pu for the reference monopile was esti-
the model pile in order to evaluate its ultimate static lateral mated as 140 N, which corresponded to a point on its
load-carrying capacity, with and without the piled-cruciform experimental load–rotation curve (Figure 6) where, apparently,
attachment in situ. The lateral deflection response of the the surrounding sand began to yield substantially; that is, its
pilehead was monitored using two horizontally mounted dis- rotation increased by 0·4° over the load increment from 120 to
placement transducers (Figure 4), with the transducer readings 140 N, whereas it increased by almost 1·0° for the load incre-
allowed to stabilise before the application of the next load ment from 140 to 160 N. For this point (140 N), the model
increment. The rotation response of the rigid pile was calcu- pile had rotated by 1·5° from its initial vertical alignment, pro-
lated using these displacement measurements, as described in ducing a lateral deflection (measured at the sand bed surface
the paper by Arshad and O’Kelly (2014). level) of  7 mm; that is, 0·13  D. From Figure 6, it can be
interpreted that the inclusion of the different piled-cruciform
Different assumptions have been used by researchers regarding arrangements all substantially improved the ultimate static
the determination of the ultimate static lateral load-carrying lateral load-carrying capacity of the monopile, as summarised
capacity, although they are generally based on excessive lateral in Table 3 for quick comparison.
displacement of the pilehead or rotation of the pile (Hu et al.,
2006; Nasr, 2014; Peng et al., 2011). Some researchers deter-
7. Performance of piled-cruciform
mine the value of Pu as corresponding to a point on the load–
arrangement in reducing monopile
deflection (rotation) curve where the pile starts to deflect
rotation under cyclic lateral loading
(rotate) significantly for a relatively small increase in the lateral
load (Dickin and Laman, 2003; Prasad and Chari, 1999). 7.1 Effect of SDP embedment length
Twelve tests were performed to investigate the effect of the
Figure 6 shows the static lateral load–rotation relationships SDP embedment length on the rotation response of the mono-
obtained for monopiles having five different piled-cruciform pile, considering four different loading scenarios and two
arrangements (set-ups A–E in Table 2) investigated as part of LSDP/L values of 0·5 and 0·33 (set-ups A and B, respectively,
the present study, along with reference data for the monopile in Table 2). Figures 7(a) and 7(b) show the rotation responses
alone. Previous studies of rigid, model pile behaviour per- of the monopile for one-way and partial one-way loading,
formed at 1g usually estimated the Pu value for lateral pile respectively, while Figures 7(c) and 7(d) show its responses
deflections of 0·1–0·2  D (Cuéllar et al., 2012; El Sawwaf, for balanced and unbalanced two-way loading, respectively.

220

200

180

160

140
Static load: N

120

100

80 Reference monopile
60 Set-up A
Set-up B
40 Set-up C
Set-up D
20
Set-up E
0
0 0·25 0·50 0·75 1·00 1·25 1·50 1·75 2·00
Rotation: degrees

Figure 6. Static lateral load plotted against rotation relationships


for the different experimental set-ups shown in Table 2

8
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

From these figures, it can be interpreted that the piled-cruci-


Set-up Lateral Rotation: form attachment to the monopile considerably reduced
(refer to Table 2) resistance: N degrees its accumulated rotation, when compared with the reference
monopile for the same loading condition. For instance, at 6000
Reference 90 0·5
load cycles for the unbalanced two-way loading scenario of
A 200 0·5
2w/30–60/0·40 (Figure 7(d)), the reference monopile had
B 136 0·5
rotated by 2·75°, whereas with the piled-cruciform attachment
C 122 0·5
in situ, its rotation was limited to 0·49° and 0·38° for LSDP/L
D 106 0·5
values of 0·33 and 0·5, respectively; that is, 82% and 86%
E 96 0·5
reductions in the monopile rotation were achieved (as deter-
Reference 150 1·88
mined using Equation 2). The respective reductions in the
A 150 0·32
monopile rotation were 57% and 68% for balanced two-way
B 150 0·62
loading (i.e. 2w/60–60/0·40 in Figure 7(c)), and 24% and 31%
C 150 0·76
D 150 0·96 for one-way loading (i.e. 1w/60/0·25 in Figure 7(a)). For partial
E 150 1·2 one-way loading (i.e. 1w/30–60/0·25 in Figure 7(b)), when
compared with the reference monopile, the rotation was found
Table 3. Lateral resistance mobilised for 0·5° rotation of monopile, to reduce by 73% for both LSDP/L values investigated. The
and monopile rotation produced by an applied lateral load of reduction in the monopile rotation can be largely explained by
150 N, for the different set-ups of the piled-cruiform attachment the additional passive resistances mobilised along the

1·4 0·7
1w/30–60/0·25
1·2 0·6
1w/30–60/0·25/A
1·0 0·5 1w/30–60/0·25/B
Rotation: degrees
Rotation: degrees

0·8 0·4

0·6 0·3

0·4 1w/60/0·25 0·2


1w/60/0·25/A
0·2 0·1
1w/60/0·25/B
0 0
0 2000 4000 6000 8000 10 000 12 000 14 000 0 2000 4000 6000 8000 10 000 12 000 14 000
Number of load cycles Number of load cycles
(a) (b)

1·8 3·0
2w/60–60/0·40 2w/30–60/0·40
1·6
2w/60–60/0·40/A 2·5 2w/30–60/0·40/A
1·4
2w/30–60/0·40/B
Rotation: degrees
Rotation: degrees

2w/60–60/0·40/B
1·2 2·0
1·0
1·5
0·8
0·6 1·0
0·4
0·5
0·2
0 0
0 2000 4000 6000 8000 10 000 12 000 14 000 0 2000 4000 6000 8000 10 000 12 000 14 000
Number of load cycles Number of load cycles
(c) (d)

Figure 7. Effect of SDP embedment length on accumulated


rotation of the monopile: (a) one-way loading; (b) partial one-way
loading; (c) balanced two-way loading (d) unbalanced two-way
loading

9
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

embedded lengths of the four SDPs, as described in Section 2. 125·0 to 88·4 mm) between the SDP centres and the axis of the
monopile orthogonal to the loading direction (see Figure 10),
 
ab which may have the effect of causing some greater overlap
2: Reduction ðas %Þ ¼  100
a between the zones of significant influence for the laterally loaded

where a is the monopile rotation and b is the rotation of the


monopile with the piled-cruciform attachment in situ. 1·2

Figure 8 shows the percentage reduction in the monopile 1·0

Rotation: degrees
rotation produced when the LSDP/L ratio was increased from
0·33 to 0·5. From this figure, it can be concluded that, for this 0·8
50% increase in the LSDP/L value, the monopile rotation was
0·6
reduced by between 9% and 27% over the 12 000 load cycles
applied. Hence, for practical purposes, based on these model 0·4
test results, there is no major benefit achieved in increasing the
LSDP/L ratio from 0·33 to 0·5. 0·2 1w/60/0·25/B
1w/60/0·25/C
0
7.2 Effect of loading direction orientation 0 2000 4000 6000 8000 10 000 12 000 14 000
Six of the tests performed were designed to investigate two differ- Number of load cycles
ent orientations of the piled-cruciform arrangement, with respect (a)
to the line of action of the applied loading (set-ups B and C in
0·8
Table 2), considering three different loading scenarios and an
LSDP/L value of 0·33. Figure 9 shows that relatively greater rota- 0·7
tion of the monopile occurred when the cruciform arms were
Rotation: degrees

0·6
offset by 45° from the line of action of the applied loading (set-
0·5
up C), as compared with full alignment (set-up B), although the
difference was marginal for the one-way and balanced two-way 0·4
loading scenarios investigated. For instance, at 6000 load cycles, 0·3
the monopile had rotated by 0·92° (1w/60/0·25/C in Figure 9(a))
0·2
and 0·73° (2w/60–60/0·40/C in Figure 9(b)) for set-up C, com-
0·1 2w/60–60/0·40/B
pared with 0·88° and 0·69°, respectively, for set-up B. However,
2w/60–60/0·40/C
the effect was significant for unbalanced two-way loading 0
(2w/30–60/0·4 in Figure 9(c)), with monopile rotations of 0·76° 0 2000 4000 6000 8000 10 000 12 000 14 000
and 0·49° measured for set-ups C and B, respectively, at 6000 Number of load cycles
load cycles. A possible reason for the greater rotation experienced (b)
0·9
for set-up C may be the decrease in perpendicular distance (from
0·8
0·7
Rotation: degrees

30
0·6

25 0·5
Reduction in rotation: %

0·4
20
0·3
15 0·2
2w/30–60/0·40/B
0·1
10 2w/30–60/0·40/C
0
1w/60/0·25 0 2000 4000 6000 8000 10 000 12 000 14 000
5 2w/30–60/0·40 Number of load cycles
2w/60–60/0·40 (c)
0
0 2000 4000 6000 8000 10 000 12 000 14 000
Number of load cycles Figure 9. Effect of orientation of piled-cruciform attachment,
with respect to line of action of the applied loading, on
Figure 8. Percentage reduction in accumulated rotation of accumulated rotation of monopile: (a) one-way loading;
monopile achieved by increasing LSDP/L value from 0·33 to 0·5 (b) balanced two-way loading; (c) unbalanced two-way loading

10
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

0·8
0·7
0·6

125 mm

Rotation: degrees
88·4 mm
0·5
0·4
0·3
0·2
Monopile 2w/60–60/0·40/B
0·1
2w/60–60/0·40/D
SDPs 0
0 2000 4000 6000 8000 10 000 12 000 14 000
Number of load cycles
(a)

1·0
0·9
Figure 10. Relative positions of SDPs when the cruciform arms
0·8
were aligned with, and offset by 45° from, the line of action of
Rotation: degrees

0·7
the applied loading
0·6
0·5
monopile and SDPs. For practical reasons, it can be considered 0·4
that the more favourable orientation occurs for the cruciform 0·3 2w/30–60/0·40/B
arms aligned with the line of action of the applied loading. 0·2
2w/30–60/0·40/D
0·1
2w/30–60/0·40/E
0
7.3 Effect of SDP inclination 0 2000 4000 6000 8000 10 000 12 000 14 000
The efficiency of the piled-cruciform arrangement was also eval- Number of load cycles
uated for different inclinations (to the vertical direction) of the (b)
SDPs at an LSDP/L value of 0·33. For this purpose, the four
1·2
SDPs were inclined outward, at 15° to the vertical direction, and
the monopile tested under balanced and unbalanced two-way 1·0
loading, with the cruciform arms aligned with the line of action
of the applied loading (set-up D in Table 2). Figures 11(a) and 0·8
Rotation: degrees

11(b) show the effect of this change in inclination, comparing


set-up D with set-up B (see Table 2). From these figures, it 0·6
can be concluded that, for both of these loading scenarios inves-
tigated, the set-up with the inclined SDPs produced greater 0·4
2w/30–60/0·40/SDPs inclined at 15°
monopile rotation. For instance, compared with the vertical
SDPs (set-up B), the monopile rotation for the SDPs inclined at 0·2 2w/30–60/0·40/SDPs inclined at 25°
15° under balanced two-way loading (2w/60–60/0·4/D in 2w/30–60/0·40/SDPs inclined at 40°
0
Figure 11(a)) was 5–9% greater between the 1000th and 6000th 0 200 400 600 800 1000 1200 1400 1600 1800 2000
load cycles. For unbalanced two-way loading (2w/30–60/0·4/D Number of load cycles
in Figure 11(b)), the corresponding increase was in the range (c)
16–25% (the same percentage range increase also occurring for
up to the maximum number of load cycles of 12 000 applied). Figure 11. Effect of SDP inclination to vertical direction on
accumulated rotation of monopile: (a) 15°, balanced two-way
Another test was performed for unbalanced two-way loading, loading; (b) 15°, unbalanced two-way loading; (c) 15°, 25° and
with the SDPs inclined at 15°, but with the cruciform arms 40°, unbalanced two-way loading
offset by 45° from the line of action of the applied loading
(set-up E in Table 2). Compared with set-up B (i.e. cruciform
arms aligned with the lateral loading direction), the monopile A possible reason for the increase in monopile rotation for
rotation for this arrangement (2w/30–60/0·4/E in Figure 11(b)) the inclined SDPs may be their lower resistance against the
was 97% greater at the 1000th load cycle, reducing to 74% at pulling action of the applied lateral loading, which is certainly
the 12 000th load cycle. greater for vertical SDPs. This hypothesis was investigated by

11
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

performing some additional tests, each involving 1500 cycles 2000


of unbalanced two-way lateral loading (2w/30–60/0·4), for the 1800

Cyclic stiffness: N/mm


same LSDP/L value of 0·33, with the cruciform arms in line 1600
with the loading direction, but setting the inclination of the 1400
SDPs at 15°, 25° and 40° to the vertical direction (Figure 11(c)). 1200
1000 1w/60/0·25/B
At the 1500th load cycle, the monopile rotation values for the 1w/60/0·25/C
800
25° and 40° settings were found to be 17% and 28% greater 1w/60/0·25
600
compared with the 15° setting. It is the authors’ view that it is 1w/30–60/0·25/A
400
probable this situation could be reversed if these SDPs were 1w/30–60/0·25/B
200 1w/30–60/0·25
replaced by helical piles of similar dimensions. 0
0 2000 4000 6000 8000 10 000 12 000
8. Effect of piled-cruciform on cyclic Number of load cycles
stiffness of soil–pile system (a)
The cyclic stiffness of the soil–pile system was calculated and
600
plotted in Figure 12 for the series of tests reported in Tables 1
and 2. The overall trend in the variation of cyclic stiffness with 500
increasing number of load cycles was broadly similar for set-
Cyclic stiffness: N/mm

ups with and without the piled-cruciform attached to the 400


monopile. For a given set-up, a marginal overall increase in
cyclic stiffness occurred for one-way loading (Figure 12(a)); for 300
example, compared with the reference monopile (1w/60/0·25),
200 2w/60–60/0·40/A
the values of cyclic stiffness mobilised with the piled-cruicform
2w/60–60/0·40/B
attachment in situ (set-ups B and C) were typically 20–30% 100 2w/60–60/0·40/C
greater over the 1000–6000 load cycle range. For partial one- 2w/60–60/0·40
way loading (1w/30–60/0·25 in Figure 12(a)), the increase in 0
0 2000 4000 6000 8000 10 000 12 000
cyclic stiffness was significantly greater (typically 70–130%
Number of load cycles
over the same load cycle range for set-ups A and B).
(b)
For balanced two-way loading with the piled-cruciform attach- 600
ment in situ (set-ups A–C in Figure 12(b)), the values of cyclic
stiffness increased from  230 N/mm to 380 N/mm over the 500
Cyclic stiffness: N/mm

1000–6000 load cycle range (and to 450 N/mm at 10 000 load


400
cycles), whereas the reference data increased from  170 to
265 N/mm over the same range. For unbalanced two-way 300 2w/30–60/0·40/A
loading over the 1000–6000 load cycle range, with the piled- 2w/30–60/0·40/B
cruciform attachment in situ (considering set-ups A–C in 200 2w/30–60/0·40/C
Figure 12(c)), the cyclic stiffness increased from  350 to 2w/30–60/0·40/D
100 2w/30–60/0·40/E
415 N/mm, compared with  195 to 310 N/mm for the refer-
2w/30–60/0·40
ence data (2w/30–60/0·4). Comparing the responses for set-ups 0
A and B, it was found that the initial values and rates of 0 2000 4000 6000 8000 10 000 12 000
increase in cyclic stiffness were marginally greater for set-up A, Number of load cycles
which can be explained by its greater LSDP/L value of 0·5 (c)
(compared with 0·33 for set-up B), although overall, the differ-
ence was not great. This is consistent with the earlier finding Figure 12. Cyclic stiffness response of the different foundation
that there was no major benefit (in terms of the reduction in set-ups: (a) one-way loading; (b) balanced two-way loading; (c)
monopile rotation) achieved in increasing the LSDP/L ratio unbalanced two-way loading
from 0·33 to 0·5 (see Figures 7 and 8).

Achmus et al. (2009), API (2010) and DNV (2011) have sug- soil next to the monopile. The present experimental findings
gested that, irrespective of soil type, the stiffness of the soil– also indicate an increase in foundation stiffness with increasing
pile system degrades under cyclic lateral loading. In contrast, number of lateral load cycles.
Rosquoët et al. (2007), LeBlanc et al. (2010), Bhattacharya
et al. (2011) and Cuéllar et al. (2012) have reported that the The change in stiffness of the soil–pile system with increasing
foundation stiffness actually increases with the number of number of load cycles may adversely affect the performance of
lateral load cycles, on account of densification of the sandy the structure supported by the monopile. For instance, an

12
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

increase in cyclic stiffness of the foundation system would 10. Conclusions


increase its natural frequency of vibration, which could poten- The improvement in performance produced by the piled-
tially lead to resonance occurring when the forcing frequency cruciform attachment to the monopile head was investigated
and the natural vibration frequency of the system come close to for many thousands of lateral load cycles. Based on the results
one another. This would cause increased deflection (rotation) of the different series of tests presented, the following con-
of the monopile, which in turn may cause more rapid deterio- clusions can be drawn.
ration of the on-board machinery and, in some cases, may ulti-
mately lead to structural failure (Adhikari and Bhattacharya, (a) The accumulated rotation of the monopile was dependent
2011). In the case of a strain-hardening site (e.g. loose to on the cyclic load characteristics, the embedment depth
medium-dense sand or normally consolidated clay deposit), the and inclination of the SDPs and the loading direction
natural vibration frequency of the foundation system is expec- orientation (i.e. orientation of the cruciform arms relative
ted to increase, whereas for a strain-softening site (e.g. dense to the line of action of the applied loading).
sand or overconsolidated clay deposit), its natural frequency of (b) For four vertical SDPs having LSDP/L values of 0·33 and
vibration would decrease (Bhattacharya et al., 2011, 2013). 0·5, the greater reduction in monopile rotation at 6000 load
cycles was recorded for unbalanced two-way loading (82%
9. Limitations of this study and 86%, respectively), followed by partial one-way loading
The present investigation was performed using a model pile, (73% for both), balanced two-way loading (57% and 68%,
partially embedded in dense sand beds (density index range of respectively) and finally one-way loading (24% and 31%,
70–74%). Additional testing should be undertaken to evaluate respectively). The performance of the piled-cruciform
the behaviour of the monopile for other density ranges and modification reduced when the cruciform arms were offset
different soils types. Further, the field stress conditions can be by 45° from the line of action of the applied loading.
simulated more adequately within the soil mass (beds) using a (c) The monopile rotation reduced for greater embedment of
centrifuge testing facility, rather than performing the tests at 1g. the SDPs, although for the LSDP/L values of 0·33 and 0·5
investigated, comparable overall reductions in the monopile
It may be a limitation of this study that only 6000 load cycles rotation were achieved for the experimental conditions
were applied in some of the tests performed, because the num- investigated. Hence, the shorter SDPs (i.e. one-third of the
ber of load cycles considered for the fatigue limit state of an monopile length) may be adequate for practical purposes.
offshore wind-turbine foundation is significantly greater (d) Vertical SDPs were more efficient in reducing the
(Germanischer Lloyd, 2005). It was found that the general accumulated rotation of the monopile compared with
trend in the monopile behaviour under cyclic lateral loading SDPs inclined at 15°, 25° and 45° (the last of these was
had practically stabilised by the 6000th load cycle; that is, the least efficient among those investigated) to the vertical
although the monopile continues to rotate for greater numbers direction.
of load cycles, the experimental data plots presented indicate
that this occurs at a decreasing rate. Any quick change in the For the prototype, the cruciform arms have to be strong and
experimental curve describing the rotation against number of stiff enough to transmit the forces and the monopile will poss-
load cycles relationship, as occurred during the first 500 load ibly require some strengthening near the head level (possibly
cycles applied, seems unlikely. achieved through increased wall thickness of the monopile) to
avoid buckling.
The model pile used in this investigation geometrically
represents the field monopile at a scale of approximately Acknowledgement
1:100. Among other technical difficulties encountered in per- The first author gratefully acknowledges a postgraduate
forming physical tests with geomaterials, the direct scaling research award received from Trinity College Dublin.
of the particle (grain) dimensions is particularly problematic.
Undesirable forces may be introduced unless a certain mini-
mum ratio is maintained between the effective grain size d50 REFERENCES
and a characteristic dimension of the model (pile diameter for Achmus M (2010) Design of axially and laterally loaded piles
the present investigation) (Sedran et al., 2001; Verdure et al., for the support of offshore wind energy converters.
2003). Examples of the derivation of scaling laws for monopile Proceedings of the Indian Geotechnical Conference
foundation testing at 1g have been reported by Lai (1989), GEOtrendz, Mumbai, India, pp. 92–102.
Muir Wood et al. (2002), LeBlanc et al. (2010), Bhattacharya Achmus M, Kuo YS and Abdel-Rahman K (2009) Behavior of
et al. (2011) and Cuéllar et al. (2012), although the satisfaction monopile foundations under cyclic lateral load. Computers
of such scaling conditions for similarity in granular soils, and and Geotechnics 36(5): 725–735.
scaling issues related to the stress field corresponding to hom- Adhikari S and Bhattacharya S (2011) Vibrations of wind-
ologous points, is not trivial (Bhattacharya et al., 2011; Dong turbines considering soil-structure interaction. Wind and
et al., 2001). Structures 14(2): 85–112.

13
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

API (American Petroleum Institute) (2010) API RPA2: Behaviour under Short-term and Long-term Cyclic Loading.
Recommended Practice for Planning, Designing and PhD thesis, Technical University of Berlin, Berlin,
Constructing Fixed Offshore Platforms – Working Stress Germany.
Design, 22nd edn. API, Washington, DC, USA. Cuéllar P, Georgi S, Baeßler M and Rucker W (2012) On the
Arshad M and O’Kelly BC (2013) Offshore wind-turbine quasi-static granular convective flow and sand densification
structures: a review. Proceedings of the Institution of Civil around pile foundations under cyclic lateral loading.
Engineers – Energy 166(4): 139–152. Granular Matter 14(1): 11–25.
Arshad M and O’Kelly BC (2014) Development of a rig to study Davie JR and Sutherland HB (1978) Modeling of clay uplift
model pile behaviour under repeating lateral loads. resistance. Journal of the Geotechnical Engineering
International Journal of Physical Modelling in Geotechnics Division, ASCE 104(6): 755–760.
14(3): 54–67. Dickin EA and Laman M (2003) Moment response of short
Arshad M and O’Kelly BC (2016a) Analysis and design of rectangular piers in sand. Computers and Structures
monopile foundations for offshore wind-turbine structures. 81(30–31): 2717–2729.
Marine Georesources and Geotechnology, http://dx.doi.org/ DNV (Det Norske Veritas) (2011) DNV–OS–J101: Design of
10.1080/1064119X.2015.1033070. Offshore Wind Turbine Structures. DNV, Oslo, Norway.
Arshad M and O’Kelly BC (2016b) Reducing monopile rotation Dong P, Newson TA, Davies MCR and Davies PA (2001) Scaling
under lateral loading in sandy soils. Geomechanics and laws for centrifuge modelling of soil transport by turbulent
Geoengineering, http://dx.doi.org/10.1080/17486025.2016. fluid flows. International Journal of Physical Modelling in
1153730. Geotechnics 1(1): 41–45.
Arshi HS, Stone KJL, Vaziri M et al. (2013) Modelling of Dührkop J and Grabe J (2008) Improving the lateral bearing
monopile-footing foundation system for offshore structures capacity of monopiles by welded wings. Proceedings of the
in cohesionless soils. Proceedings of the 18th International 2nd BGA International Conference on Foundations (ICOF
Conference on Soil Mechanics and Geotechnical 2008), Dundee, UK (Brown MJ, Bransby MF, Brennan AJ
Engineering, Paris, France, vol. 3, pp. 2307–2310. and Knappett JA (eds)). IHS BRE Press, Garston, UK,
Bhattacharya S and Adhikari S (2011) Experimental validation vol. 1, pp. 849–860.
of soil–structure interaction of offshore wind turbines. El Sawwaf M (2006) Lateral resistance of single pile located
Soil Dynamics and Earthquake Engineering 31(5–6): near geosynthetic reinforced slope. Geotechnical and
805–816. Geoenvironmental Engineering 132(10): 1336–1345.
Bhattacharya S, Lombardi D and Muir Wood D (2011) Similitude Germanischer Lloyd (2005) Guideline for the Certification of
relationships for physical modelling of monopile-supported Offshore Wind Turbines. Germanischer Lloyd WindEnergie
offshore wind turbines. International Journal of Physical GmbH, Hamburg, Germany.
Modelling in Geotechnics 11(2): 58–68. Haiderali A, Cilingir U and Madabhushi G (2013) Lateral and
Bhattacharya S, Cox J, Lombardi D and Muir Wood D (2013) axial capacity of monopiles for offshore wind turbines.
Dynamics of offshore wind turbines supported on two Indian Geotechnical Journal 43(3): 181–194.
foundations. Proceedings of the Institution of Civil Hu Z, McVay M, Bloomquist D, Herrera R and Lai P (2006)
Engineers – Geotechnical Engineering 166(2): 159–169, Influence of torque on lateral capacity of drilled shafts in
http://dx.doi.org/10.1680/geng.11.00015. sands. Geotechnical and Geoenvironmental Engineering
Bienen B, Dührkop J, Grabe J, Randolph MF and White D (2012) 132(4): 456–464.
Response of piles with wings to monotonic and cyclic Irvine JH, Allan PG, Clarke BG and Peng JR (2003) Improving
lateral loading in sand. Geotechnical and Geoenvironmental the lateral stability of monopile foundations.
Engineering 138(3): 364–375. Proceedings of the International Conference on Foundations:
Broms BB (1964) Lateral resistance of piles in cohesionless Innovations, Observations, Design and Practice, Dundee,
soils. Journal of the Soil Mechanics and Foundation UK (Newson TA (ed.)). Thomas Telford, London, UK,
Engineering Division, ASCE 90(SM3): 123–156. pp. 371–380.
Brown DA, Morrison C and Reese LC (1988) Lateral load Karthigeyan S, Ramakrishna VVGST and Rajagopal G
behavior of pile group in sand. Journal of Geotechnical (2006) Influence of vertical load on the lateral response
Engineering, ASCE 114(11): 1261–1276. of piles in sand. Computers and Geotechnics 33(2):
Carswell W, Arwade SR, DeGroot DJ and Lackner MA (2015) 121–131.
Soil–structure reliability of offshore wind turbine monopile Kuo YS, Achmus M and Abdel-Rahman K (2012) Minimum
foundations. Wind Energy 18(3): 483–498. embedded length of cyclic horizontally loaded monopiles.
Chari TR and Meyerhof GG (1983) Ultimate capacity of rigid Geotechnical and Geoenvironmental Engineering 138(3):
single piles under inclined loads in sand. Canadian 357–363.
Geotechnical Journal 20(4): 849–854. Lai S (1989) Similitude for shaking table test on soil–structure–
Cuéllar P (2011) Pile Foundations for Offshore Wind Turbines: fluid model in 1-g gravitational field. Soils and Foundations
Numerical and Experimental Investigations on the 29(1): 105–118.

14
Geotechnical Engineering Piled-cruciform attachment to monopile
head reduces deflection
Arshad and O’Kelly

Offprint provided courtesy of www.icevirtuallibrary.com


Author copy for personal use, not for distribution

LeBlanc C, Houlsby GT and Byrne BW (2010) Response of stiff Prasad YVSN and Chari TR (1999) Lateral capacity of model
piles in sand to long-term cyclic lateral loading. rigid piles in cohesionless soils. Soils and Foundations
Géotechnique 60(2): 79–90. 39(2): 21–29.
Lombardi D, Bhattacharya S and Muir Wood D (2013) Dynamic Rao SN, Ramakrishna VGST and Raju GB (1996) Behavior of
soil–structure interaction of monopile supported wind pile-supported dolphins in marine clay under lateral
turbines in cohesive soil. Soil Dynamics and Earthquake loading. Journal of Geotechnical Engineering, ASCE
Engineering 49(2013): 165–180. 122(8): 607–612.
Malhotra S (2010) Design and construction considerations for Rosquoët F, Thorel L, Garnier J and Canepa Y (2007) Lateral
offshore wind turbine foundations in North America. In cyclic loading of sand-installed piles. Soils and Foundations
Proceedings of GeoFlorida 2010: Advances in Analysis, 47(5): 821–832.
Modeling and Design, Orlando, FL, USA (Fratta DO, Sedran G, Stolle DFE and Horvath RG (2001) An investigation
Puppala AJ and Muhunthan B (eds)). ASCE, Reston, VA, of scaling and dimensional analysis of axially loaded piles.
USA, Geotechnical Special Publication no. 199, vol. 2, Canadian Geotechnical Journal 38(3): 530–541.
pp. 1533–1542. Stone K, Newson T and Sandon J (2007) An investigation of the
Moayed RZ, Mehdipour I and Judi A (2012) Undrained lateral performance of a ‘hybrid’ monopile-footing foundation for
behavior of short pile under combination of axial, lateral offshore structures. In Proceedings of the 6th International
and moment loading in clayey soils. Kuwait Journal of Conference on Offshore Site Investigation and Geotechnics,
Science and Engineering 39(1B): 59–78. London, UK. Society for Underwater Technology, London,
Muir Wood D, Crewe AJ and Taylor CA (2002) Shaking table UK, pp. 391–396.
testing of geotectnical models. International Journal of Terzaghi K (1955) Evaluation of coefficients of subgrade
Physical Modelling in Geotechnics 2(1): 1–13. reaction. Géotechnique 5(4): 297–326.
Nasr AMA (2014) Experimental and theoretical studies of Tomlinson MJ (2001) Foundation Design and Construction,
laterally loaded finned piles in sand. Canadian 7th edn. Pearson, Harlow, UK.
Geotechnical Journal 51(4): 381–393. Uncuoğlu E and Laman M (2011) Lateral resistance of a short
Nicolai G and Ibsen LB (2014) Small-scale testing of cyclic rigid pile in a two-layer cohesionless soil. Acta Geotechnica
laterally loaded monopiles in dense saturated sand. Journal Slovenica 8(2): 19–43.
of Ocean and Wind Energy 1(4): 240–245. Verdure L, Garnier J and Levacher D (2003) Lateral cyclic
O’Kelly BC and Arshad M (2016) Offshore wind turbine loading of single piles in sand. International Journal of
foundations – analysis and design. In Offshore Wind Physical Modelling in Geotechnics 3(3): 17–28.
Farms: Technologies, Design and Operation (Ng C and Zhang L, Silva F and Grismala R (2005) Ultimate lateral
Ran L (eds)). Woodhead Publishing, Cambridge, UK, resistance to piles in cohesionless soils. Geotechnical and
ch. 20, pp. 589–610. Geoenvironmental Engineering 131(1): 78–83.
Peng J, Clarke B and Rouainia M (2011) Increasing the Zhu B, Byrne BW and Houlsby GT (2013) Long-term lateral
resistance of piles subject to cyclic lateral loading. cyclic response of suction caisson foundations in sand.
Geotechnical and Geoenvironmental Engineering 137(10): Geotechnical and Geoenvironmental Engineering 139(1):
977–982. 73–83.

WHAT DO YOU THINK?


To discuss this paper, please email up to 500 words to the
editor at journals@ice.org.uk. Your contribution will be
forwarded to the author(s) for a reply and, if considered
appropriate by the editorial panel, will be published as
discussion in a future issue of the journal.
Proceedings journals rely entirely on contributions sent in
by civil engineering professionals, academics and stu-
dents. Papers should be 2000–5000 words long (briefing
papers should be 1000–2000 words long), with adequate
illustrations and references. You can submit your paper
online via www.icevirtuallibrary.com/content/journals,
where you will also find detailed author guidelines.

15
View publication stats

You might also like