Earthq Engng Struct Dyn - 2023 - Hamilton - Seismic Excitation of Offshore Wind Turbines and Transition Piece Response

Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Received: 13 April 2022 Revised: 23 February 2023 Accepted: 28 February 2023

DOI: 10.1002/eqe.3872

RESEARCH ARTICLE

Seismic excitation of offshore wind turbines and transition


piece response
Christopher M. Hamilton Christelle N. Abadie

Department of Engineering, University of


Cambridge, Cambridge, UK Abstract
Expansion of the offshore wind industry in seismically active areas has raised
Correspondence
Christelle N. Abadie, Department of
concerns regarding the structural integrity of offshore wind turbines under earth-
Engineering, University of Cambridge, quake loading. This paper details a 3D finite element study to investigate the
Cambridge, UK. behaviour of the structure, and in particular the transition piece (TP), under
Email: [email protected]
seismic loads. The work focuses on equivalent grouted connection and TP-less
designs, selected as promising design solutions for seismic zones. The numeri-
cal model is validated against a medium-scale 4-point bending laboratory test,
and scaled up to a representative 8 MW turbine. The results show that cracking
of the grout occurs due to earthquake excitation at the top of the TP; a location
that is typically undamaged during monotonic loading. This can lead to excessive
settlement of the transition piece and loss of axial capacity caused by deteriora-
tion of the grout-steel bond and water ingress. The residual hub displacement
after earthquake excitation is around 0.1 m. The global monotonic response post-
earthquake excitation is not significantly altered, with apparent stiffness and
ultimate strength maintained. An equivalent TP-less design is shown to have a
5% lower natural frequency than an equivalent grouted connection design, sug-
gesting a reduced global stiffness, with reduced structural damping due to the
absence of grout. However, TP-less design eliminates the risk of grout deteriora-
tion and settlement and may therefore be a safer design option for offshore wind
turbines installed in seismic zones in the future.

KEYWORDS
finite element analysis, grouted connections, offshore wind turbines, seismic, TP-less,
transition piece

1 INTRODUCTION

Offshore wind plays an important role in the transition to low carbon energy supply, and has developed rapidly over the
past few decades. It has enormous potential for providing renewable energy across the globe, and Asia Pacific waters are
foreseen to become the largest offshore wind market by 2030.1 However, many wind farms will have to be installed in
seismically active zones, and therefore the structural integrity of offshore wind turbines (OWTs) under seismic loading is
of increasing concern.

This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium, provided the
original work is properly cited.
© 2023 The Authors. Earthquake Engineering & Structural Dynamics published by John Wiley & Sons Ltd.

Earthquake Engng Struct Dyn. 2023;1–24. wileyonlinelibrary.com/journal/eqe 1


10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
2 HAMILTON and ABADIE

FIGURE 1 Schematic overview of a representative OWT model with a grouted connection (GC) and TP-less design.

80% of currently installed offshore wind turbines are installed on monopile (MP) foundations,2 with a transition piece
(TP) connecting the foundation to the superstructure (Figure 1). The most common type of MP-TP connection is a grouted
connection (GC), consisting of a cementitious ultra-high-strength grout cast between concentric steel tubes. Load transfer
occurs via shear friction mobilised by normal stress induced through interlocking of imperfections and compression of
the grout. Grouted connections were first used for oil and gas installations, with transfer of the technology to offshore
wind applications in 2002.3
However, in 2009, DNV highlighted concerns regarding insufficient axial capacity of grouted connections following
observations of significant settlement of newly installed transition pieces.4 This was attributed to the direct application of
design practice from the oil and gas industry to monopiles of much larger diameters, and hence higher flexural rigidity,
with significantly higher overturning loads. Following experimental testing, Dallyn et al.4 concluded that the significantly
larger bending moments observed in the grouted connection in offshore wind structures (compared with oil and gas instal-
lations) resulted in ovalisation of the tubular connection and the formation of gaps between the grout and the surrounding
steel in the connection. This led to a reduced axial capacity of the connection and excessive relative displacements under
loads significantly lower than the ultimate limit state. The effect can also be worsened by the possibility of water ingress
between the grout and steel following the formation of cracks in the grout, further reducing the integrity of the bond and
axial capacity. This led to adjustments in the DNV-OS-J101 guidelines3 (now known as DNVGL-ST-01265 ) and improved
design practice, including the use of conical interface surfaces3 and the addition of bolts.6 The most commonly adopted
modification consists of the addition of circumferential welds to act as shear keys in the grout (Figure 1) as this signifi-
cantly reduces settlement while retaining manufacturing simplicity.3,7 The design guidelines consider fatigue damage in
grouted connections due to cyclic loading under operational loads,5 but reduced confidence in grouted connections in
the offshore wind industry remains and the effect of large stress cycles on the connection due to seismic loads is not well
documented.
Alternative designs for the connection have been explored and adopted, and Table 1 provides a comparison of these
alternatives. Bolted flange connections commonly use M72 bolts and have been deployed since 2009.8 However, this type
of connection is vulnerable to loosening of nuts9 under dynamic loads, and therefore may be a higher risk solution in
seismic areas. Slip joint connections have been developed much more recently,10,11 and deployed for the first time in 2020.12
They consist of a simple friction connection between conical surfaces of the monopile and transition piece, with integrity
of the connection introduced through vibration-assisted gravity fitting of the transition piece. However, settlement of the
connection during installation has been observed for frequencies below 20 Hz,10 and therefore harmonic excitation of the
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
HAMILTON and ABADIE 3

TA B L E 1 Comparison of current MP-TP connection design practices.

Installation Manufacture Expected


Connection Manufacturing Installation Installation Corrosion damage installation seismic
type complexity complexity speed risk sensitivity Size limit tolerances OPEX CAPEX performance
Grouted Low High (grout Medium Low Low Vessel Large High Low Unknown
prep) (historically)
Bolted Medium High Low High High 8-10MW Small Medium Medium Poor
Flange (internal (<M72)
flange)
Slip Joint High (conical Low High Low Medium Vessel Medium Low High Medium
interface)
TP-less High High High Low High (WTG Much Medium Low Medium Good
flange) larger
vessel
required

connection under high-frequency seismic loading could potentially cause disengagement of the friction connection and
excessive displacements.
The most recent design development is to construct the monopile and tower as a single piece (TP-less, Figure 1). Ini-
tial assessment of this type of design by Empire Engineering and Wood Thilsted13 estimates a reduction in operational
expenditure (OPEX) of 6%–10%, and opportunity for significant life extension. This design method is only now becoming
feasible due to the introduction of larger installation vessels which can transport considerably longer monopiles. There is
significant risk associated with the increased cost of manufacture of the support structure as one piece, and the necessity
of driving the pile directly on the tower interface flange. As a result, widespread deployment of TP-less designs is likely to
take time. Nevertheless, the simplicity of the design is likely to offer good performance under seismic loads.
It is likely that designers will pursue designs with grouted connections in seismically active zones in the immedi-
ate future, particularly due to the accumulated experience of using this type of connection in the industry, and design
improvement. TP-less designs may offer a good alternative once the technology and supply chain are mature.
Research of the structural behaviour of offshore wind turbines under dynamic loads is increasing. De Risi et al.14 Vacare-
anu et al.15 and Kaynia16,17 showed high sensitivity of OWT structures to historical seismic events, accounting for the
combination of seismic and environmental (wind/wave) loading. Sensitivity to the operational conditions (e.g. idle or
maximum power generation) was also observed by Mo et al.,18 highlighting the importance of considering operational
loads in seismic analyses. However, these studies did not model the transition piece explicitly.
This paper focuses on establishing a 3D finite element model to assess the seismic response of the transition piece and
support structure of a representative 8 MW turbine with both a grouted connection and TP-less design (Figure 1). Valida-
tion of the numerical model is achieved via comparison with a medium-scale experimental test on a grouted connection
under quasi-static four-point bending, performed by Wilke.19 The specific interest of this study is to verify the gapping
failure mode in the grouted connection identified by Dallyn et al.20 under seismic excitation and assess cracking, and
to investigate the effect of the connection on the dynamic response of the structure. This paper provides a comparison
of the performance of a grouted connection compared with a TP-less design under seismic excitation, and assesses the
post-seismic monotonic response of both structures, including latent hub displacements, stiffness and strength.

2 NUMERICAL MODELLING

The results presented in this paper were obtained through 3D finite element analysis using the general-purpose mul-
tiphysics finite element package LS-DYNA.21 LS-DYNA has been specifically designed for dynamic analyses with high
degrees of non-linearity, and therefore offers excellent explicit time integration capability. It has formulations for advanced
contact modelling, and includes robust constitutive models developed to capture cracking in cementitious materials.
Monotonic analyses were performed using the implicit solver, enabling higher computational efficiency over the explicit
solver for static loading conditions. Dynamic analyses were performed using implicit–explicit switching, with an initial
implicit analysis to preload the structure, followed by dynamic excitation using the explicit solver, and finally followed
by an implicit analysis to obtain the post-earthquake monotonic response. This numerical modelling scheme was chosen
following work by Asnaashari et al.22 and Vieira et al.23 who demonstrated the effectiveness of 3D FE analysis to model
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
4 HAMILTON and ABADIE

grouted connections under static loads. Natural frequencies and modeshapes were determined using the Block Shift and
Invert Lanczos method.21

2.1 Geometry and discretisation

Dynamic analyses were performed using 2 full-scale models: one with a grouted connection, and one with a TP-less design.
Geometries were chosen to capture a representative design for the support structure of an 8 MW turbine based in South-
East Asia (e.g. Formosa 1), based on the Vestas V164 8 MW wind turbine generator.24 A monopile diameter of 8.75 m was
selected, both because it is representative of an 8 MW turbine support structure, and also because this corresponds to one
of the standard design cases of the PISA project25 (Geometry D225,26 ), allowing for the use of a representative PISA soil
response (Section 2.3). A water depth of 30 m was chosen, representative of installations in East Asia. The dimensions of
the upper structure were then selected to be coherent with the foundation, and based on typical dimensions to support
the Vestas V164 8 MW turbine generator. The key dimensions of the models are provided in Figure 1.

2.1.1 Grouted connection model

The dimensions of the grouted connection are in accordance with DNVGL-ST-0126 guidelines5 (Figure 1), with shear keys
occupying the centre third of the grouted length. Exact dimensions were adjusted to be coherent with the medium-scale
laboratory model detailed by Wilke,19 whose test data was also used for model validation (Section 3). The height of the TP
flange above the mean sea level was 22.9 m, with mean sea level positioned at the centre of the grouted connection.
The support structure was discretised using eight-node hexahedral solid elements with a constant stress formula-
tion, with the exception of the central elements of each shear key, which used six-node tetrahedral solid elements. Solid
elements were selected over shell elements to better capture through-thickness stresses and to allow for more robust fric-
tional contact to be defined between the steel and grout. A minimum of three elements through the material thickness
was ensured throughout the model. The mesh of the GC model was refined around the connection itself, as shown in
Figures 2A and 2B. Table 2 provides further details on the dimensions of the connection used in the GC model.

2.1.2 TP-less model

The TP-less support structure geometry consists of a linear taper from the monopile diameter at the mudline (𝐷𝑀𝑃 =
8.75 m), to the diameter of the base of the turbine tower (𝐷𝑡𝑜𝑤𝑒𝑟 = 7.7 m), with the same wall thickness as the monopile
(𝑡𝑀𝑃 = 91 mm). No localised mesh refinement was required for the TP-less model (Figure 2C).

F I G U R E 2 (A) Grouted connection model with zoomed view of mesh refinement around the shear keys; (B) cross-sectional view;
(C) TP-less model.
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
HAMILTON and ABADIE 5

TA B L E 2 Dimensions of the grouted connection, compared with DNVGL-ST-0126 guidelines.5 .


Dimension Value DNV limit
MP diameter, 𝐷𝑝 (mm) 8750 10 ≤ 𝑅𝑝 ∕𝑡𝑝 ≤ 30
MP thickness, 𝑡𝑝 (mm) 91 n/a
TP diameter, 𝐷𝑇𝑃 (mm) 9114 9 ≤ 𝑅𝑇𝑃 ∕𝑡𝑇𝑃 ≤ 70
TP thickness, 𝑡𝑇𝑃 (mm) 91 n/a
Grout thickness, 𝑡𝑔 (mm) 150 n/a
Grout length, 𝐿𝑔 (mm) 13125 n/a
Overlap length, 𝐿𝑔 ∕𝐷𝑝 1.5 1.5 ≤ 𝐿𝑔 ∕𝐷𝑝 ≤ 2.5
Number of shear keys, 𝑛 11 n/a

Shear key spacing, 𝑠 (mm) 600 𝑠> 𝑅𝑝 𝑡𝑝
Shear key height, ℎ (mm) 30 ℎ ≥ 5 𝑚𝑚, ℎ ≤ 0.1 𝑠
Shear key width, 𝑤 (mm) 60 1.5 ≤ 𝑤∕ℎ ≤ 3.0

TA B L E 3 Dimensions of the tower used in the GC and TP-less models (adapted from LEANWIND 8MW reference turbine27 ).
Elevation above TP flange (m) Can diameter (mm) Wall thickness (mm)
0.0 7700 50
15.0 7320 50 → 45
27.5 7000 45 → 40
37.5 6740 40 → 35
47.5 6490 35 → 30
55.0 6300 30 → 28
62.5 6110 28 → 27
80.0 5660 27 → 26
90.0 5410 26 → 25
106.0 5000 25

2.1.3 Superstructure

The LEANWIND 8MW reference turbine geometry27 was used to model the turbine tower, with the exception of the
individual tower can thicknesses. Early monotonic analyses of the structure exhibited unrealistic local buckling of the
tower base under relatively small loads. Therefore, the diameters and heights of each can were taken from the reference
geometry (Table 3), but the thicknesses were adjusted to obtain sufficient bending stiffness and reasonable bending stress
distribution, while maintaining representative natural frequencies of the structure. A similar average thickness over the
length of the tower was maintained compared with the reference turbine. The entire turbine assembly (including blades)
was lumped at hub height as a 495 tonne point mass, in-line with the axis of the tower. Many turbine models have an
offset of the rotor-nacelle assembly centre of gravity from the tower axis, but this distance is small compared with the
vertical distance between the TP and the hub (i.e., the tower height), which dominates the bending moment in the tower
and foundation. As such, this offset and its effect on the three-direction mass inertia of the RNA was not considered. The
turbine tower was discretised using fully integrated quadrilateral shell elements with a significantly coarser mesh density
than for the support structure, as the global structural response of the tower was of interest, rather than local stresses. The
tower was connected to the top of the transition piece using rigid connections between adjacent nodes, as this connection
is suitably far from the grouted connection, and is usually a bolted flange allowing limited relative displacements in the
plane of the section. Table 3 details the tower can dimensions, with arrows indicating a step-change in wall thickness at a
particular elevation.
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
6 HAMILTON and ABADIE

TA B L E 4 Material properties for S355 steel and grout (after DIN EN 10225-428 and Wilke19 ).
Property S355 Steel Grout
Density, 𝜌 (kg/m3 ) 7850 2380
Poisson’s ratio 0.3 0.3
Elastic modulus, 𝐸 (MPa) 210503 50000
Yield strength, 𝑓𝑦 (MPa) 355.5 n/a
Tangent modulus, 𝐸𝑡𝑎𝑛 (MPa) 1097 n/a
Compressive strength, 𝑓𝑐 (MPa) n/a 130
Tensile strength, 𝑓𝑡 (MPa) n/a 7
Fracture energy, 𝐺𝑓 (Nm/m2 ) n/a 150.8

FIGURE 3 Shear cracking mode under compression captured by the Winfrith model.

2.2 Material models

2.2.1 Steel

A bilinear elastic–plastic material model with isotropic hardening was used for all steel components. S355 steel is com-
monly used for OWT structures, and the material properties defined according to DIN EN 10225-428 were used here
(Table 4).

2.2.2 Grout

The physical properties of the grout were defined according to Wilke19 and Ducorit S5 product specifications (the grout
manufacturer),29 and are representative of the grout used for grouted connections offshore.
The Winfrith material model was used to model the grout. This is a plasticity model that simulates Mohr-Coulomb like
behaviour, with the inclusion of the third stress invariant that enables both triaxial compression and triaxial extension to
be accounted for.30 The model is based on the four parameter model proposed by Ottosen,31 and allows for up to three
orthogonal crack planes per element. Figure 3 shows a schematic view of a block of elements failing in the shear cracking
mode under compressive load captured by the material model. A fracture energy approach was used32 to avoid using a
strain formulation, which has been reported to cause numerical instabilities in concrete-related analyses.33 The fracture
energy, 𝐺𝐹 was determined from the CEB model code for concrete structures34 :

( )0.7
𝑓cm
𝐺𝐹 = 𝐺𝐹0 − (1)
𝑓cm0

where 𝐺𝐹0 is the base fracture energy as a function of aggregate size, and 𝑓𝑐𝑚0 = 10 MPa. The material properties for the
grout used in the FE analysis are summarised in Table 4.
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
HAMILTON and ABADIE 7

2.3 Soil-structure interaction

The response of the foundation to lateral loads was captured using lateral and rotational macro-element springs located
at the mudline (Figure 1). This was chosen as an alternative to the more traditional 1-D p-y curve method due to the abil-
ity to capture load reversal behaviour under cyclic loading. The load-displacement behaviour of the soil under repeated
load cycles, as well as the resulting hysteretic damping, has a significant effect on the dynamic response of the structure,
and needs to be accounted for in seismic design. While not providing information about the detailed load distribu-
tion along the pile, a macro-element model allows for a very efficient representation of the soil structure interaction
for structural analysis. This efficiency was required in this case to permit more detailed and complex analysis of the
transition piece. A discussion of the advantages of this modelling technique, compared with 1-D representations, is
provided by Abadie et al.,35,36 showing the computational efficiency of a well-calibrated macro-element model similar
to the one used in this paper. This approach also allows for simple application of the seismic loads, with the ground
motions applied to the free end of each spring (Section 2.5.2). The lateral and rotational macro-element springs were
assumed to be uncoupled, which is unlikely to be the case in reality and is a limitation of this work. Further research
is needed to develop a rigourous formulation of this coupling, and there is currently very limited guidance in published
literature.
The unload-reload response needed to capture the response to cyclic loading was obtained through a kinematic hard-
ening model. The hysteretic loop follows the Masing rule,37 with the response to perfectly symmetric loading such that (i)
the unloading and reloading curves are defined based on the initial loading curve (the backbone curve) scaled by a factor
of 2 and (ii) after any load reversal, the tangent shear modulus is the same as the initial shear modulus of the backbone
curve. This captures a reverse loading response that is representative of the macro response of the pile, but relies greatly on
an appropriate choice of backbone curve, discussed in the following sections. The model does not account for ratcheting,
gapping and excess pore water pressure35,38 and was implemented through a parallel-series Iwan model.39,40 Further work
is needed in the future to account for these effects, with excess pore water pressure and local liquefaction along the pile
shaft being of particular interest. However, there is not currently any robust spring model able to capture this adequately
for monopile design.

2.3.1 Lateral response

The backbone curve for the torsional 𝑀 − 𝜓𝐺 and lateral 𝐻 − 𝑣𝐺 springs were chosen based on the PISA work,25,26 with 𝑀
and 𝐻 being the moment and lateral load at ground level, respectively, and 𝜓𝐺 and 𝑣𝐺 the pile rotation and displacement
at ground level, respectively. The PISA project was a joint industry project run through the Carbon Trust from 2013 to
2017.26 The project enabled the development of improved numerical methods for the design of monopiles subjected to
monotonic loading. The findings are now widely adopted by the offshore industry, and it is regarded as a robust method
for the prediction of the response of large diameter monopiles at ground level. PISA 2 (2017-2018)25 expanded the PISA
design method to layered soil profiles. In this paper, the case of a relatively realistic stiff layered soil profile (E3) was
chosen, consisting of alternating layers of Cowden till and Dunkirk sand of varying relative densities (Figure 4A). The
macro-response published by Burd et al.25 was used to calibrate the backbone curve, using the 𝑀 − 𝜓 (Figure 4A) 𝐻 − 𝑣𝐺
(Figure 4B) response of an 8.75 m diameter monopile with an embedment of 35 m.
Figure 4B also shows the 𝐻 − 𝑣𝐺 response of the final macro-element Iwan model used in the model presented here.
This is compared with the monotonic response published by Burd et al.25 for validation. Particular focus is drawn to the
small displacement region of the curve, as accurate estimation of the small–strain stiffness plays an important role in the
determination of the structural natural frequencies.

2.3.2 Axial response

The axial response was not considered in the PISA work, and instead was derived from first principles and the API design
method.41 The backbone curve response was idealised as elastic-perfectly plastic. The ultimate axial capacity was derived
using the API method for driven piles in sand and clay,41 with the shaft resistance integrated over each soil layer. A
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
8 HAMILTON and ABADIE

F I G U R E 4 (A) Selected stiff soil profile, and moment-rotation response at the mudline; (B) Lateral force-displacement response of the
monopile at the mudline in the stiff soil profile.

capacity of 185 MN was obtained for the selected soil profile (Figure 4A). The axial pile stiffness was derived based on
the assumption of a rigid pile42 :

𝑉 4𝑅𝑏𝑎𝑠𝑒 𝐺𝑏𝑎𝑠𝑒 2𝜋𝐿𝐺𝑎𝑣𝑔


= + (2)
𝑤𝐺 1−𝜈 4

where 𝑉 is the vertical load and 𝑤𝐺 is the pile vertical settlement at ground level (1). 𝑅𝑏𝑎𝑠𝑒 is the pile radius (4.35 m). 𝐺𝑏𝑎𝑠𝑒
is the shear stiffness at the pile base, 𝐺𝑎𝑣𝑔 is the mean shear stiffness along the pile shaft, and 𝜈 is the Poisson’s ratio. This
elasto-plastic model was also implemented through an Iwan formulation to enable modelling of load reversal following
the Masing rule.37

2.4 Boundary conditions

Seismic excitation was applied in all three directions at ground level, and therefore no axis of symmetry could be used
and the entire structure was modelled. The nodes of the monopile at the mudline were connected by a nodal rigid body
(NRB), allowing global displacement and rotation of the node set, but no relative displacements within the set. The SSI
macro-element springs described in Section 2.3 were connected to the central node of the mudline NRB. In addition to
this SSI condition, a clamped support condition was also analysed for comparison by constraining this mudline NRB in
all 6 degrees of freedom. Going forward, the two support conditions are referred to as ‘stiff soil’ and ‘clamped’.
Contact was defined between the grout and steel surfaces using a penalty-based formulation and a coefficient of friction
of 𝜇 = 0.4.
A non-reflecting boundary was defined at the mudline using an impedance matching formulation in order to avoid
contamination of the results of the dynamic analyses by the re-entry of reflected stress waves at the fixed end of the
soil springs.

2.5 Loading

2.5.1 Quasi-static loads

Load arising from self-weight was simulated using translational base acceleration in the vertical direction equal to
9.81 m s−2 , resulting in the appropriate inertial body force loads being imposed on the nodes of the structure due to gravity.
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
HAMILTON and ABADIE 9

The importance of considering operational loads when considering the seismic response has been highlighted by
Esfeh43 and Kaynia.16 For this reason, operational wind and wave loads were applied to the structure during all dynamic
analyses. To allow for independent assessment of the dynamic response to seismic excitation alone, the operational loads
were considered to be quasi-static, and the dynamic excitation due to wind and wave loads was not considered. The quasi-
static loads were based on an operating state of maximum power generation at a wind speed at hub height 𝑣ℎ𝑢𝑏 = 13 m s−1
and a representative sea state of significant wave height 𝐻𝑆 = 4 m and period 𝑇 = 8 s. Equivalent static loads for these
environmental conditions were calculated using the method summarised in Appendix A. The wind load on the rotor-
nacelle assembly was concentrated as a point load at hub height and was equal to 1.36 MN. The wind load on the rest
of the structure varied nonlinearly from 0.06 kN m−1 at mean sea level to 0.2 kN m−1 at hub height. Wave load varied
nonlinearly from 143 kN m−1 at mean sea level to 48 kN m−1 at the mudline. The directions of wind and wave loading
were assumed to be coincident, to capture a severe loading condition.

2.5.2 Seismic loads

Ground motions were applied to the model as imposed velocities at the free end of each soil spring, with baseline correction
to eliminate drift. Applying ground motions as velocities is a common practice to avoid instability and error associated
with double integration of acceleration time histories when imposing displacement instead. The most severe direction of
ground motion for the historical earthquake records considered was aligned with the directions of both wind and wave
loading, to capture the most severe case. Ground motions were applied in all three directions using available accelerometer
time histories.44
Given the complexity of the grouted connection model and the associated high computational cost, the selection of a
limited number of relevant earthquake motions was essential. As such, two earthquake records were selected for this study,
enabling direct comparison with published work on offshore turbines under seismic loads17,18,45 : (i) Kobe, Japan, Nishi-
Akashi, January 17, 1995, and (ii) Chi-Chi Taiwan-03, CHY082, September 20, 1999. These were also selected because of
their significantly contrasting characteristics, enabling further insight into the dynamic behaviour of the structure under
varying seismic loads:
∙ The Kobe motion has a high peak ground acceleration (PGA), of 0.5 g, whereas the CHY082 Chi-Chi motion has a much
lower PGA of 0.05 g
∙ Ground displacements are higher during the Chi-Chi motion compared to the Kobe motion
∙ The two records have very different frequency characteristics – the Kobe motion is largely composed of higher fre-
quency accelerations, with the major peak at approximately 2 Hz, whereas the CHY082 Chi-Chi motion has a more
even distribution of accelerations, with significant low frequency oscillation between 0.1 and 1 Hz

The ground motions observed during the Chi-Chi earthquake were heavily dependent on local geological conditions,
with high PGA and frequencies observed south of the epicentre, and low PGA and very low frequency (and subsequent
high ground displacements) in the north.46 The CHY082 station observed the latter characteristics, and was chosen in this
paper as a comparison to the Kobe motion. Figures 5A and 5B show the unscaled ground acceleration and displacement
time histories for the two earthquake records, in the most severe accelerometer direction.
Finally, in addition to the historical input motions, an artificial sinusoidal input of magnitude 0.3𝑔 at the first natural
frequency 𝑓0 of the structure was also analysed, with an acceleration of:
𝑎 = 0.3𝑔 sin (2𝜋𝑓0 𝑡) (3)
The sinusoidal excitation was applied in one direction only, aligned with the direction of operational loading, allowing
for more rigourous investigation of the structural response at resonance.
Seismic analyses were followed by a monotonic lateral load, applied at the hub height and up to the capacity of the
structure, allowing for investigation of the post-seismic performance of the structure.

2.6 Damping

The overall damping of an offshore wind turbine consists of aerodynamic, hydrodynamic, structural, and soil damp-
ing. Aerodynamic damping was estimated to be 3.5% from Liu et al.,47 for a wind speed of 13 m s−1 . Hydrodynamic
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
10 HAMILTON and ABADIE

F I G U R E 5 Ground motions selected for dynamic analysis in the worst-case accelerometer direction: (A) Ground acceleration and
(B) ground displacement time histories.

damping consists of radiation and viscous damping, and values suggested by Germanischer Lloyd48 of 0.22% and 0.15%
were used. Structural damping was estimated based on material damping of steel of 0.3%.48 Soil damping was not
added, both because it is already accounted for by the energy contained within the hysteresis loop of the constitu-
tive model described in Section 2.3, and also because it is difficult to estimate accurately.48 The influence of additional
damping through the use of tuned mass dampers in the tower was not considered, but could form part of future
research.
Damping was applied using frequency-independent damping stresses applied to the elements of the structure, a
technique developed by Richard Sturt and Yuli Huang (Arup) for use in automotive and seismic vibration prob-
lems, and implemented in LS-DYNA.21 The damping model allows for uniform damping to be applied to a wide
frequency range, does not damp rigid body modes which would not be dissipative in reality, and captures dimin-
ished damping once stresses exceed the elastic range.49 Structural damping was applied to all elements of the model,
and aerodynamic and hydrodynamic damping was applied only to those elements above or below the water line,
respectively.

2.7 Analysis programme

The numerical analysis programme involved (i) validation of the model, (ii) eigenvalue analysis, (iii) seismic analysis,
and (iv) monotonic analysis, both before and after seismic excitation. Validation was achieved via comparison with
experimental test data for a medium-scale model, described in Section 3.
The seismic analyses consumed the vast majority of the computational time, and dictated the choice of combinations
of geometry, support condition and input motion to analyse. All combinations of support condition and geometry were
investigated for the Kobe ground motion, as it is the most representative record for earthquakes in East Asia. The stiff soil
support condition for both geometries was then used for analysis with the Chi-Chi ground motion. Both the grouted
connection and TP-less models with stiff soil support conditions were subjected to sinusoidal oscillation at the first
natural frequency.
Eigenvalue analyses and pre- and post-seismic monotonic analyses were performed for all combinations of geometry
and support conditions.
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
HAMILTON and ABADIE 11

3 VALIDATION

3.1 Medium-scale physical test model

Very little published data on physical tests is available to validate numerical models of transition pieces, largely due to the
size of the tests required. No data was found regarding dynamic response, but Wilke19 performed a quasi-static 4-point
bending test on a medium-scale grouted connection which was used for validation of the material models at the same
scale. Wilke’s 4-point bending test reproduces a representative bending moment in the grouted connection to reproduce
the governing failure mode identified by Dallyn et al.4 While axial load was not considered in Wilke’s test, the test provides
very useful data for validation of the numerical model as it is capturing the behaviour under the governing load in isolation.
A schematic of this test is shown in Figure 6A. Figure 6B shows the equivalent 3D FE model developed for validation,
making use of the symmetry plane to reduce computational cost. The simply-supported boundary conditions were repli-
cated through the use of point constraints. The grout material used in the test model (Ducorit S5) was representative of
grout used in full-scale offshore installations, particularly with regards to geometric specifications (e.g., aggregate size),
and as such is representative of the behaviour of the grout at full-scale. The frictional capacity of a full-scale connection
would be expected to be lower than that of the small scale test specimen because of (i) the scale independency of fabrica-
tion parameters, typically resulting in a relatively rougher surface for smaller diameters; and (ii) the reduction in activated
hoop/compression stress and associated reduced bond strength with increasing diameter. However, Wilke19 quantified this
reduction for a range of connection designs, showing that this reduction is very small for grouted connections with shear
keys, where the response is dominated by the compressive strength of the grout and is effectively independent of scale
effects. As a result, this test model captures the behaviour of a full-scale grouted connection well.
The steel material properties used in the FE model were defined according to the tensile tests performed by Wilke.19
Load was applied to a set of nodes on each loading flange over an area equivalent to that of the physical test. The
load eccentricity present in the physical test was replicated through calculation of the equivalent forces applied to each
loading flange.
The FE model was validated by (i) comparing the behaviour of the model under 4-point bending with the experimental
results obtained by Wilke19 and (ii) comparing the settlement of the transition piece onto the pile during 4-point bend-
ing with the settlement predicted by the analytical model presented by Lotsberg et al.50 (Section 3.2), and later included
in a revised version of the DNV code for the design of wind turbine support structures, DNV-OS-J101, now known as

F I G U R E 6 Medium-scale physical test: (A) Wilke test set-up19 ; (B)Corresponding 3D FE analysis model for validation with zoomed
view of mesh refinement around the shear keys.
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
12 HAMILTON and ABADIE

FIGURE 7 Validation of numerical model: (A) Displacements at point 1 (Figure 6B); (B) Gap formation at point 2 (Figure 6B).

DNVGL-ST-0126.5 Tziavos et al.51 also developed a detailed 3D FE model of this test specimen to study the effect of
additional shear keys on the settlement of the transition piece, and these numerical results were used for further validation.
The response of the FE model during the 4-point bending test is assessed by plotting the displacement of reference
point 1 (Figure 6B) in the direction of loading. The results shown in Figure 7A demonstrate very good agreement between
the FE models and the experimental data, with the initial stiffness, peak displacement and latent plastic deformation
within 2% of the experimental results. The gap opening at reference point 2 (Figure 6B) also shows good agreement, with
the maximum gap opening within 6% of the experimental results (Figure 7B). The stress distribution and magnitudes
observed in the FE model also displayed good agreement with both the numerical and experimental results from Wilke
and Tziavos, respectively. These results provide confidence in the material models selected for the numerical analysis,
along with their constitutive properties.

3.2 Comparison with analytical model

The analytical model from Lotsberg et al.50 was used as a secondary approach to validate the FE model. The total moment
at transition piece level, 𝑀𝑡𝑜𝑡 is related to the nominal radial contact pressure, 𝑝 at the top and bottom of the connection:

3𝜋𝑀𝑡𝑜𝑡 𝐸𝐿𝑔
𝑝= ( ) (4)
𝑅𝑝2 2
𝑅𝑇𝑃
𝐸𝐿𝑔 {𝑅𝑝 𝐿𝑔2 (𝜋 + 3𝜇) + 3𝜋𝜇𝑅𝑝2 𝐿𝑔 } + 18𝜋2 𝑘𝑒𝑓𝑓 𝑅𝑝3 +
𝑡𝑝 𝑡𝑇𝑃

where 𝑅𝑝 , 𝑡𝑝 , 𝑅𝑇𝑃 and 𝑡𝑇𝑃 are the radius and thickness of the monopile and transition piece, respectively. 𝐸 is the Young’s
modulus of the steel, 𝐿𝑔 is the grout length and 𝜇 is the friction coefficient of the grout-steel interface. 𝑘𝑒𝑓𝑓 is an effective
vertical stiffness per circumferential length representing the resistance of the shear keys, derived based on structural
mechanics as:
2
2𝑡𝑇𝑃 𝑠𝑒𝑓𝑓 𝑛𝐸
𝑘𝑒𝑓𝑓 = (5)
⎧( ) 3 ⎫
√ ⎪ 𝑅 2 ( )3 ⎪
𝑅 2
4 3(1 − 𝜈2 )𝑡𝑔2 ⎨
4 𝑝 2
+ 𝑇𝑃 ⎬𝑡𝑇𝑃 + 𝑛𝑠𝑒𝑓𝑓 𝐿𝑔
𝑡 𝑡𝑇𝑃
⎪ 𝑝 ⎪
⎩ ⎭
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
HAMILTON and ABADIE 13

F I G U R E 8 Relative displacement of the transition piece (settlement onto the pile) vs. applied load at reference point 2 (Figure 6B):
(A) Medium-scale test19 ; (B) Large scale geometry, representative of an 8 MW OWT.

where 𝑛 and 𝑠𝑒𝑓𝑓 are the effective number and spacing of the shear keys, 𝑡𝑔 is the grout thickness, and 𝜈 is the Poisson’s ratio
of steel. The contact pressure is then related to the vertical displacement of the transition piece relative to the monopile
𝛿𝑣 :
( )
6𝑝 𝑅𝑝 𝑅𝑝2 𝑅2
𝛿𝑣 = + 𝑇𝑃 (6)
𝐸 𝐿𝑔 𝑡𝑝 𝑡𝑇𝑃

Figure 8A shows the displacement predicted by the analytical model at the final applied load of 1 MN for the model
test described in Section 3.1. This is also compared with the results from Tziavos.7 There is very good agreement between
the FE model and the analytical model, particularly at the final applied load of 1 MN, providing further confidence in the
numerical modelling of the frictional interface between the grout and the steel.
As the validated model is of a smaller scale than the full-scale models described in Section 2 and used for subsequent
dynamic analysis, the test geometry was scaled up by a factor of 10. This enabled the modelling of an 8 m diameter grouted
connection, with the applied load increased to 100 MN to achieve a similar level of plasticity. This provided a geometry
similar to the geometry of the grouted connection model (Section 2.1), to bridge the gap between the medium-scale and
full-scale models. Scaling of the grout material model is not necessary, as the grout used in the experimental test was
representative of grout used for full-scale installations in terms of strength and aggregate size. Scaling of the frictional
characteristics between the grout and steel is also not necessary, as explained in Section 3.1 and shown by Wilke.19
Figure 8B shows the displacements predicted by the analytical and FE models at this larger scale. There is very good
agreement at the final applied load of 100 MN, providing further confidence in the model described in Section 2, and
confidence in the extension of the model to full-scale.

4 GROUTED CONNECTION RESPONSE TO SEISMIC EXCITATION

4.1 Mechanism of deformation

Grouted connections have been shown to be susceptible to the formation of gaps between the grout and steel due to
ovalisation of the circular cross-section under large bending moments.20 Figures 9A to 9E clearly show this mechanism
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
14 HAMILTON and ABADIE

F I G U R E 9 Plan and elevation cross-sections through the grouted connection with displacements magnified 500x: (A) undeformed;
(B) under operational load only; (C) 10 s after the start of the Kobe ground motion, at the peak of the first major oscillation of the structure in
its first bending mode (see point 1 on Figure-12); (D) after the Kobe ground motion has finished; (E) under 4x operational load after seismic
excitation.

through the course of the dynamic analysis, with displacements magnified 500x for clarity. Under operational load alone
the connection ovalises subtly, and gaps form at the top and bottom of the connection. The deformed shape agrees well
with that observed by Wilke,19 with a characteristic s-shaped deformation, more pronounced in the less stiff pile section.
As the earthquake progresses and the structure oscillates in its first bending mode, the ovalisation of the circular cross-
section significantly worsens, the gaps increase in size and the grout begins to debond from the steel in these areas. After
the ground motion has finished, the ovalised shape returns to that observed before excitation, but the grout-steel bond is
compromised due to the repeated cycles of severe ovalisation.This is analysed further in Section 4.2.
Finally, Figure 9E shows the deformed shape under a monotonic load equivalent to four times the operational load. This
is applied after seismic excitation has finished, capturing an extreme monotonic load to ultimate capacity. Unsurprisingly,
this extreme case exhibits more ovalisation of the connection than the other cases, and a similar deformed shape would be
obtained without prior earthquake excitation. Nevertheless, the ovalised shape observed under maximum bending during
the earthquake (Figure 9C) is similar to that observed in Figure 9E, highlighting the severity of the deformation caused
by seismic excitation.

4.2 Grout deterioration

Along with gap formation, damage of the grout material itself can lead to water ingress as the grout fails and powder forms,
which can result in loss of bond extent, corrosion and excessive settlement.20 Therefore, confidence in grout integrity
during and after seismic excitation is essential. The load transfer mechanism through the grout occurs through the devel-
opment of compressive struts between adjacent shear keys, clearly evidenced by the numerical analyses. Figure 10 shows
contours of the compressive stress in the grout at the centre line cross section on the primarily compressive side of the
connection (opposite operational load application – point 2 in Figure 1) 15 s after the start of the Kobe ground motion.
During dynamic oscillation of the structure these compressive struts undergo corresponding oscillation in magnitude,
and on the opposite side of the connection similar oscillating tensile zones occur in the grout.
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
HAMILTON and ABADIE 15

FIGURE 10 Contours of compressive stress in the grout at the centre line cross section on the compressive side at 𝑡 = 15𝑠.

FIGURE 11 Damage in the grout on the compressive side: (A) pre-earthquake, under operational load only; (B) after the Kobe ground
motion.

However, during the earthquake and subsequent free vibration, the cyclic loading of the grout between adjacent shear
keys causes significant cracking. Figures 11A and 11B show damage in the grout before and after the earthquake. In these
Figures, blue indicates an uncracked, healthy element, green indicates initialisation of a crack, with strains on the soft-
ening curve (tensile stresses remain), light orange indicates a developed but closed crack (compressive stresses remain),
and magenta indicates a fully cracked element (the crack is open and no tensile/compressive stress can be generated).
Under operational load (Figure 11A), the initialisation of cracks remains limited, and is confined to the upper shear
keys on the compressive side of the connection. On the tensile side, the development of cracks is uniform across all shear
keys, but limited in extent to approximately 10% of the circumference for each shear key. This failure mode of the grout is
in good agreement with experimental observations from Wilke19 and numerical computations from Tziavos et al.51
Following the earthquake (Figure 11B), the circumferential cracks on the tensile side developed and propagated to
approximately 50% of the circumference at each shear key location. Cracks on the compressive side at the top of the
connection also developed more severely. These zones of extensive damage are due to initiation and propagation of cracks
in in-plane shear (Mode II, Figure 3), caused by the large compressive stress concentration at the top of the transition
piece during dynamic oscillation of the structure.
Figure 12 shows a time history of percentage of cracked elements in this compressive region in the top 1.5 m of the
connection during seismic excitation, alongside the lateral TP displacement (Figure 13A Point 2) and input acceleration
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
16 HAMILTON and ABADIE

FIGURE 12 Percentage of cracked elements in the top 1.5 m of the grouted connection during the Kobe ground motion.

time histories. It is clear that the initiation and propagation of cracks (shown by the accumulation of cracked elements)
correspond with the large oscillations of the structure in the first mode, with significant propagation after each compres-
sive cycle. By the end of the motion, approximately 8% of the elements in the top 1.5 m of the connection have cracked,
and these elements are almost exclusively on the compressive side. Under monotonic loading alone, even at four times the
operational load, no cracking in the top 1.5 m of the connection was observed. This cracking is likely to reduce the grout-
steel bond strength as grout powder formation occurs, and significantly increase the risk of water ingress and corrosion.
Further cyclic loading could cause progressive failure further down the connection as the extent of bond is reduced, and
further impact capacity. It is of particular concern that while the structure is expected and designed to be able to withstand
an earthquake event like the one analysed, significant damage may be caused to the grout that could significantly impact
its lifespan, and this is not something that is currently considered in design, with cracking in the upper compressive zone
unique to seismic excitation. Design of connections for future use in seismic zones should require remediation of the high
stress concentrations at the top and bottom of the connection, as well as consideration of cracking between adjacent shear
keys to avoid risk of extensive grout damage and unexpected settlement and reduction in capacity.

5 EFFECT OF TRANSITION PIECE ON SEISMIC RESPONSE

5.1 Modal analysis and natural frequencies

Natural frequencies and modeshapes for each model were determined using the implicit solver and the Block Shift and
Invert Lanczos method in LS-DYNA. Due to the symmetry of the structure, identical modes exist in both the fore-aft and
side to side directions, and for the purpose of this modal analysis one of these directions was constrained. Only bending
modes were considered, so the rotational degree of freedom about the axis of the monopile was also constrained. Figure 13A
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
HAMILTON and ABADIE 17

F I G U R E 1 3 (A) Modal response of the grouted connection model: First three bending modes; (B) Normalised spectral acceleration for
the 8MW LEANWIND turbine comparing the effect of the transition piece on the modal response.

shows exaggerated views of the first, second and third modeshapes, with reference points labelled for analysis in the
following sections. The support type of transition piece (grouted connection or TP-less) were not found to significantly
alter the mode shapes.
Figure 13B shows the normalised peak spectral acceleration for the 8 MW LEANWIND reference turbine, with the com-
puted natural frequencies and earthquake ground motions. This is shown on a logarithmic scale to facilitate comparison
with the earthquake spectra. The first natural frequency in all cases was found to lie between the 1P (turbine rotation)
and 3P (blade passing) frequency ranges (the ‘soft-stiff’ domain), which is in line with design practice. Figure 13B shows
the effect of the transition piece on the natural frequencies of the structure. While the equivalent TP-less structure has
less steel at connection level, and therefore a reduced global stiffness, the inclusion of the (softer) grout also softens the
response. Figure 13B shows that the first natural frequency for the TP-less model supported by stiff soil is approximately
5% lower than that of the grouted connection model, with the same conclusion for the clamped support condition. The
second natural frequency is very similar for both models (Figure 13A).

5.2 Hub displacement during seismic excitation

Figure 14 shows a comparison between the relative hub displacement due to the Kobe ground motion for the grouted
connection and TP-less models. The reduced stiffness of the TP-less model is clearly evidenced by the lower frequency of
oscillation. For the stiff soil case, displacements in the first 10 s of excitation are almost identical between the two models,
but the second peak is approximately 5% larger for the TP-less geometry. This amplification is magnified further for the
case with a clamped support condition, with peak hub displacements 20% larger with the TP-less geometry.
Figure 14 shows approximately 0.4% greater structural damping in the case of the grouted connection model with a
clamped support condition, with the grout contributing to the overall structural damping.
Latent deformation at hub height after earthquake excitation can also be seen for both geometries, due to soil failure
and associated deformation at the mudline. While the hub displacements observed during this severe dynamic excita-
tion are temporary and may be acceptable for the turbine due to being accounted for with generator shutdown and
other fault avoidance controller strategies, any permanent deformation due to soil failure is of particular importance
due to the often very tight out-of-vertical alignment tolerances that the turbine generator can withstand during normal
operation.
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
18 HAMILTON and ABADIE

FIGURE 14 Effect of connection type on hub displacements in the direction of operational load during the Kobe ground motion.

5.3 Response to low-frequency seismic excitation

Figure 15A shows hub displacement during the Chi-Chi ground motion. The peak displacement observed for the TP-less
geometry is approximately 20% larger than for the grouted connection. However, the peak displacements for both geome-
tries are 40%–60% larger than for the Kobe ground motion, despite the peak ground acceleration of the Chi-Chi earthquake
being an order of magnitude lower. Accelerations at hub height are also high, with amplification factors exceeding 2. This
highlights the severe impact of the low frequency components of the Chi-Chi motion. The dynamic response in all cases
is dominated by the first mode, and even small ground accelerations near the first natural frequency can cause severe
excitation of this mode.
The GC and TP-less models were subjected to six cycles of sinusoidal acceleration of magnitude 0.3𝑔 and frequency
equal to their respective first natural frequencies, to further investigate pure excitation of the first mode. Figure 15B shows
a comparison of the relative hub displacements and accelerations caused by this ground motion. Increased damping and
reduced peak displacements for the grouted connection model are also evidenced here, with the magnitudes of displace-
ment greatly increased because of the severe motion applied. While this regular sinusoidal motion at the first natural
frequency is unlikely to occur in reality, it highlights the severe hub displacements that can occur if there are frequency
components close to the natural frequency. High accelerations at hub height are again clear, with amplification factors
exceeding 2.
In both cases latent deformation at hub height is again observed, with associated risk for turbine operation as
previously discussed.
It should be noted that the above conclusions do not account for the addition of dampers in the tower of the turbine,
which would likely be added. However, these results show that the TP-less designs may be more vulnerable to seismic
excitation and may require further additional damping.

5.4 Post-seismic monotonic response

The structure was allowed to undergo free vibration with zero ground motion after each seismic event until it returned
to equilibrium. An implicit switch was then executed to perform a static analysis on the structure. The static loads were
removed, and the structure was linearly reloaded to approximately 4 times the operational load. An independent and
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
HAMILTON and ABADIE 19

F I G U R E 1 5 Effect of connection type on hub displacements in the direction of operational load: (A)Chi-Chi ground motion; (B) 0.3𝑔
sinusoidal excitation at the first natural frequency.

identical static analysis was performed for each model without prior seismic excitation, to allow for comparison between
the pre- and post-seismic response.
Figure 16A compares the load-displacement response at hub height before and after the Kobe ground motion under
varying support conditions, alongside a zoomed view of the initial section to allow for direct comparison of stiffness.
These results show that despite accumulated deformation caused by the earthquake, neither the load-deflection response
nor the global stiffness are significantly affected by the seismic event. The post-earthquake monotonic responses rejoin
the pre-seismic backbone curve after approximately 2 m of hub displacement.
It is important to note that there are limitations to this conclusion, however. The constitutive model used for the soil
response, which is a kinematic hardening model (Section 2.3), does not account for the effects of ratcheting, gapping,
excess pore water pressure and potential liquefaction on the permanent deformation and stiffness of the soil during seis-
mic excitation. These effects could in reality cause more disturbance to the monotonic response observed in Figure 16A.
Further research into the effect of excess pore water pressure and liquefaction will be needed to extend this research in
the future.43,45
The analytical model introduced by Lotsberg et al.50 to compute the vertical displacement of the transition piece relative
to the monopile under bending (Equations 4 to 6) was compared with the displacements obtained using the numerical
model. The application of the model to the experimental tests performed by Wilke19 neglected axial load, and therefore
Equation 6 was modified to account for self-weight:
( )
6𝑝 𝑅𝑝 𝑅𝑝2 2
𝑅𝑇𝑃 𝑃𝑑
𝛿𝑣 = + + (7)
𝐸 𝐿𝑔 𝑡𝑝 𝑡𝑇𝑃 2𝜋𝑅𝑝 𝑘𝑒𝑓𝑓
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
20 HAMILTON and ABADIE

F I G U R E 1 6 (A) Pre and post-seismic monotonic response; (B) Vertical displacement of the transition piece relative to the monopile,
compared with analytical prediction from Lotsberg et al.50

where 𝑃𝑑 is the self-weight of the structure above the monopile, including the full self-weight of the transition piece. The
analytical model describes a linear relationship between the relative vertical displacement and the total moment applied at
transition piece level, with an initial offset due to the self-weight. This analytical relationship is shown in Figure 16B for the
grouted connection geometry described in Section 2.1, compared with the relative displacement obtained numerically both
before and after the Kobe and Chi-Chi earthquakes, and at a total moment corresponding to roughly twice the operational
load considered in this study. Because the self-weight and monotonic loads are applied linearly for the numerical models,
comparison with the analytical model is only possible once the full self-weight has been applied, and therefore only the
final displacement is shown for the numerical models.
Very good agreement can be seen between the analytical model and the numerical model with clamped support con-
ditions. Accounting for SSI with the stiff soil response increases the relative displacement by approximately 16%, to a
maximum displacement of 2 mm. A further increase in relative displacement is observed following the Kobe earthquake
- almost 30% greater than the analytical prediction.
The analytical model is therefore suited if both the soil-structure interaction and deterioration of the grout during a
cyclic loading event are disregarded. The design of the shear keys relies on the above analytical model, and as a result the
underestimation of settlements (and corresponding shear forces on individual shear keys) shown in Figure 16B suggests
that the design calculations presented in DNVGL-ST-01265 may be under-conservative in the case of seismic loads.

6 CONCLUSIONS

This paper investigates the response of a grouted connection and equivalent TP-less design for a generic monopile-
supported 8 MW turbine design to seismic excitation. The results highlight the effects of dynamic loads on the structural
integrity and capacity of grouted connections and form a basis for further research to inform the design process. For this
purpose, this paper details a numerical finite element model for modal and transient dynamic analysis of the structure,
with detailed study of the grouted connection response to seismic excitation.
A finite element model replicating the medium-scale test performed by Wilke19 was developed for validation, and excel-
lent agreement is obtained between the numerical model and the experimental data. Further comparison with a similar
numerical study performed by Tziavos et al.51 and the analytical model used in the DNV design guidelines3,50 provides
confidence in the numerical method adopted in this paper. The model was then scaled up to simulate an 8 m diameter
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
HAMILTON and ABADIE 21

grouted connection, representative of a full-scale 8 MW offshore wind turbine, and compared with the analytical model;
the only available model available for comparison at such large scale. Excellent agreement between the numerical and
analytical models provides confidence in the scaling method.
Based on this validation, 3D finite element models of a representative 8 MW offshore wind turbine with a grouted
connection and TP-less design have been developed. Soil-structure interaction is implemented through a macro-element
model at the mudline with kinematic hardening and soil response defined according to the PISA project.25 An operational
loading condition is captured for all analyses using a quasi-static representation of the wind and wave load, disregarding
dynamic effects to allow for assessment of the dynamic response to seismic excitation in isolation. Future research extend-
ing the findings of this paper to consider the combined dynamic effects of wind and wave alongside seismic excitation
would be valuable.
Detailed analysis of the grouted connection under seismic excitation shows that a short, severe seismic event can cause
significant cracking of the grout in locations typically not observed under monotonic loading. This is of particular concern
due to historical issues of settlement and reduced capacity of grouted connections, caused by grout deterioration and
subsequent water ingress. The failure mode observed, with extensive cracking in the top region of the connection, is unique
to seismic excitation and not currently considered in design. This does not invalidate the use of grouted connections for
seismic zones, but further understanding and consideration of this behaviour under dynamic cyclic loading is required in
design. In particular, damage due to seismic loading should be considered in combination with prior accumulated fatigue
damage during normal operation, which will further add to the risk of capacity reduction.
A reduction of 5% of the natural frequency is found from an equivalent TP-less design. The presence of the grout also
increases global damping by approximately 0.4% compared with the equivalent TP-less structure. Consideration of the
reduced damping of a TP-less design and appropriate compensation if necessary should be made in practice. Excitation by
the Chi-Chi ground motion highlights that earthquakes with prominent low frequency characteristics can be catastrophic,
even at low accelerations.
Comparison with the analytical model from DNV3,50 to predict the vertical displacement of the transition piece relative
to the monopile demonstrates that the design guidelines under-predict the settlement due to lack of consideration of the
soil response. The results also show that the model, developed without accounting for the effects of seismic excitation, may
be non-conservative in the event of an earthquake during the lifetime of the turbine, potentially resulting in an increase
in relative displacement of over 30% compared with that predicted by the analytical model. Further investigation of this
behaviour and revision of the analytical model for use in seismic zones could be beneficial to ensure robust design of the
shear key arrangement in grouted connections and provide greater confidence in the axial capacity and longevity of the
connection. This will be particularly important in the next few years, while TP-less designs are still in development.

AC K N OW L E D G M E N T S
The authors are thankful for the contributions of the specialist structures and mechanical group at Arup in providing sup-
port, access to computing facilities, as well as the LS-DYNA and Oasys Suite licenses for the development of this work. The
authors also acknowledge the contribution of the PISA project team in providing data for the lateral soil response. Fruitful
advice and discussion with Alasdair Parkes (Arup), Dr Nikolaos Tziavos (University of Cambridge), Geert Weymeis (Jan
de Nul) and Professor Ali Mehmanparast (University of Strathclyde) was also very valuable for the development of this
project, together with the help of Carlos Español-Espinel and Professor Robin Langley (University of Cambridge) during
the writing phase of this paper.

D A T A AVA I L A B I L I T Y S T A T E M E N T
The data that support the findings of this study are available from the corresponding author upon reasonable request.

ORCID
Christopher M. Hamilton https://orcid.org/0000-0002-2118-3745
Christelle N. Abadie https://orcid.org/0000-0002-5586-6560

REFERENCES
1. Shuxin Lim P. Communications Manager GWECGA. Market to Watch: Asia Pacific to Become Largest Offshore Wind Market by 2030.
https://bit.ly/2TmcSWH (last accessed April 2022); 2020.
2. Wind Europe (previously EWEA). Offshore Wind Energy: 2021 mid-year statistics. https://proceedings.windeurope.org (last accessed April
2022); 2021.
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
22 HAMILTON and ABADIE

3. Lotsberg I. Developments of grouted connections in monopile foundations. In: Proceedings of the 2010 IABSE Seminar; 2010; Copenhagen,
Denmark.
4. Dallyn P, El-Hamalawi A, Palmeri A, Knight R, Morris A. Wear in large diameter grouted connections for offshore wind energy converters.
In: Proccedings of the 10th International Conference on Advances in Steel Concrete Composite and Hybrid Structures. 2012; Singapore.
5. DNV. DNVGL-ST-0126: Support structures for wind turbines. 2016. https://dnv.com/energy/standards-guidelines/dnv-st-0126-support-
structures-for-wind-turbines.html
6. Schaffer W, Seo J, Choi E, Lee J. Optimized retrofit design of in-service monopile foundation offshore wind turbine transition zone. Eng
Struct. 2020;220(111001). doi: 10.1016/j.engstruct.2020.111001
7. Tziavos NI, Hemida H, Metje N, Baniotopoulos C. Grouted connections on offshore wind turbines: a review. Proc Inst Civ Eng: comp mech.
2016;169(4):183-195. doi: 10.1680/jencm.16.00004
8. Redondo R, Mehmanparast A. Numerical analysis of stress distribution in offshore wind turbine M72 bolted connections. Met. 2020;10(689).
doi: 10.3390/met10050689
9. Ishimura M, Sawa T, Karami A, Nagao T. Bolt-nut loosening in bolted flange connections under repeated bending moments. In:
Proceedings of the ASME 2010 Pressure Vessels & Piping Division / K-PVP Conference. 2010; Washington, USA.
10. Segeren M, Lourens EM, Tsouvalas A, Zee T. Investigation of a slip joint connection between the monopile and the tower of an offshore
wind turbine. IET Renew Power Gener. 2014;8. doi: 10.1049/iet-rpg.2013.0163
11. Durakovic A. Slip Joint Foundation Makes Debut Off the Netherlands. https://bit.ly/34hoUCS (last accessed April 2022); 2020.
12. Van Oord. After years of preparation, Van Oord successfully installed the innovative world’s first submerged Slip Joint. https://bit.ly/
3vl1hFv (last accessed April 2022); 2020.
13. Empire Engineering. Foundation Ex: The tech sessions: designing the Monopile without a transition piece: why, how and when? https://
bit.ly/3fhLeCH (last accessed April 2022); 2020.
14. De Risi R, Bhattacharya S, Goda K. Seismic performance assessment of monopile-supported offshore wind turbines using unscaled natural
earthquake records. Soil Dyn Earthq Eng. 2018;109:154-172. doi: 10.1016/j.soildyn.2018.03.015
15. Vacareanu V, Kementzetzidis V, Pisano F. 3D FE seismic analysis of a monopile-supported offshore wind turbine in a non-liquefiable soil
deposit. In: Proceedings of the 2nd International Conference on Natural Hazards & Infrastructure. 2019; National Technical University of
Athens.
16. Kaynia AM. Seismic considerations in design of offshore wind turbines. Soil Dyn Earthq Eng. 2019;124:399-407. doi: 10.1016/j.soildyn.2018.
04.038
17. Kaynia A. Effect of kinematic interaction on seismic response of offshore wind turbines on monopiles. Soil Dyn Earthq Eng & Struc Dyna.
2020;50. doi: 10.1002/eqe.3371
18. Mo R, Kang H, Li M, Zhao X. Seismic fragility analysis of monopile offshore wind turbines under different operational conditions. Energies.
2017;10:1037. doi: 10.3390/en10071037
19. Wilke F. Load Bearing Behaviour of Grouted Joints Subjected to Predominant Bending. PhD thesis. Leibniz Universität Hannover. 2014;
Hanover, Germany.
20. Dallyn P, El-Hamalawi A, Palmeri A, Knight R. Experimental testing of grouted connections for offshore substructures: a critical review.
Struct. 2015;3. doi: 10.1016/j.istruc.2015.03.005
21. LS-DYNA Keyword User’s Manual, Version 971 R13, Vol. 1 and 2. https://www.dynasupport.com/manuals (last accessed April 2022); 2021.
22. Asnaashari E, Morris A, Andrew I, Hahn W, Sinha JK. Finite element modelling and in situ modal testing of an offshore wind turbine. J
Vib Eng Technol. 2018;6. doi: 10.1007/s42417-018-0018-3
23. Vieira M, Viana M, Henriques E, Reis L. Soil interaction and grout behavior for the NREL reference monopile offshore wind turbine. J
Mar Sci Eng. 2020;8:298. doi: 10.3390/jmse8040298
24. Vestas V164 8.0MW. https://pdf.archiexpo.com/pdf/vestas/vestas-v164-80-mw/88087-134417.html (last accessed April 2022); 2011.
25. Burd HJ, Abadie CN, Byrne BW, et al. Application of the PISA design model to monopiles embedded in layered soils. Géotechnique.
2020;70(11):1067-1082. doi: 10.1680/jgeot.20.PISA.009
26. Byrne BW, Burd HJ, Zdravkovic L, et al. PISA design methods for offshore wind turbine monopiles. In: Proccedings of the Offshore
Technology Conference. 2019; Houston, Texas, USA.
27. Desmond C, Murphy J, Blonk L, Haans W. Description of an 8 MW reference wind turbine. J Phys Conf Ser. 2016:753. doi: 10.1088/1742-
6596/753/9/092013
28. Deutsches Institut für Normung (DIN) DIN EN 10225-4: Weldable structural steels for fixed offshore structures - Technical delivery conditions
- Part 4: Cold formed welded hollow sections. 2019.
29. ITW Engineered Polymers APSAalborg, Denmark. Ducorit Ultra High Performance Grout. 2017.
30. Schwer L. An Introduction to the Winfrith Concrete Model. Schwer Engineering & Consulting Services; Windsor, California, USA 2010.
31. Ottosen NS. A failure criterion for concrete. J Eng Mech. 1977;103:527-535.
32. Hillerborg A, Modéer M, Petersson P. Analysis of crack formation and crack growth in concrete by means of fracture mechanics and finite
elements. Cem Concr Res. 1976;6(6):773-781. doi: 10.1016/0008-8846(76)90007-7
33. Abdelatif A, Owen J, Hussein M. Modelling the prestress transfer in pre-tensioned concrete elements. Finite Elem Anal Des. 2015;94:47-63.
doi: 10.1016/j.finel.2014.09.007
34. CEB-FIP C. Model code for concrete structures. Bulletin D’Information. 1990.
10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
HAMILTON and ABADIE 23

35. Abadie C, Byrne B, Houlsby G, Burd H, Mcadam R, Beuckelaers W. Modelling of offshore wind monopile lifetime performance. In:
Proceedings of the 4th International Symposium on Frontiers in Offshore Geotechnics (ISFOG). 2020; Austin, Texas, USA.
36. Abadie C, Beuckelaers WJAP, et al. Modelling lifetime performance of monopile foundations for offshore wind application. J Geotechn
Geoenvironment Eng. 2023. doi: 10.1061/JGGEFK/GTENG-9833
37. Masing G. Eiganspannungen und Verfestigung beim Messing. In: Proceedings for the 2nd International Congress of Applied Mechanics.
1926.
38. Abadie CN. Cyclic Lateral Loading of Monopile Foundations in Cohesionless Soils. PhD thesis. The University of Oxford, Oxford, UK; 2015.
39. Iwan WD. A distributed-element model for hysteresis and its steady-state dynamic response. J Appl Mech. 1966;33(4):893-900. doi: 10.1115/
1.3625199
40. Iwan WD. On a class of models for the yielding behavior of continuous and composite systems. J Appl Mech. 1967;34(3):612-617. doi: 10.
1115/1.3607751
41. American Petroleum Institute Recommended Practice for Planning, Designing and Constructing Fixed Offshore Platforms - Working Stress
Design. 2000.
42. Mark R, Susan G. Offshore Geotechnical Engineering (2nd ed.). CRC Press; 2011.
43. Esfeh PK, Kaynia AM. Earthquake response of monopiles and caissons for offshore wind turbines founded in liquefiable soil. Soil Dyn
Earthq Eng. 2020;136:106213. doi: 10.1016/j.soildyn.2020.106213
44. PEER. Peer Ground Motion Database. http://ngawest2.berkeley.edu/site (last accessed April 2022).
45. Seong J, Abadie CN, Haigh S, Madabhushi G. Seismic response of offshore wind monopiles in cohesionless soils. In: Proceedings of the
4th International Symposium on Frontiers in Offshore Geotechnics (ISFOG). 2020; Austin,Texas, USA.
46. Dai J, Qi X, Tong M, Lee G. 3D temporal characteristics analyses for Chi-Chi earthquake ground motions. In: Proceedings of the 13th World
Conference on Earthquake Engineering. 2004; Vancouver, Canada.
47. Liu X, Lu C, Li G, Godbole A, Chen Y. Effects of aerodynamic damping on the tower load of offshore horizontal axis wind turbines. Appl
Energy. 2017;204:1101-1114. doi: 10.1016/j.apenergy.2017.05.024
48. Germanischer Lloyd Wind Energie. Overall Damping for Piled Offshore Support Structures, Guideline for the Certification of Offshore Wind
Turbines. 2005.
49. Huang Y, Sturt R, Willford M. A damping model for nonlinear dynamic analysis providing uniform damping over a frequency range.
Comput Struct. 2019;212:101-109. doi: 10.1016/j.compstruc.2018.10.016
50. Lotsberg I, Serednicki A, Oerlemans R, Bertnes H, Lervik A. Capacity of cylindrical shaped grouted connections with shear keys in off
shore structures. Struct Eng. 2013;91:42-48.
51. Tziavos NI, Hemida H, Metje N, Baniotopoulos C. Non-linear finite element analysis of grouted connections for offshore monopile wind
turbines. Ocean Eng. 2018;171:633-645. doi: 10.1016/j.oceaneng.2018.11.005
52. ASCE, American Society of Civil Engineering ASCE/SEI 07/-10: Minimum Design Loads for Buildings and Other Structures. 2013.
53. Frohboese P, Schmuck C. Thrust coefficients used for estimation of wake effects for fatigue load calculation. In: Proceedings of the
European Wind Energy Conference. 2010; Warsaw, Poland.
54. DNV. DNVGL-ST-0437: Loads and site conditions for wind turbines. 2016. https://dnv.com/energy/standards-guidelines/dnv-st-0437-loads-
and-site-conditions-for-wind-turbines.html

How to cite this article: Hamilton CM, Abadie CN. Seismic excitation of offshore wind turbines and transition
piece response. Earthquake Engng Struct Dyn. 2023;1-24. https://doi.org/10.1002/eqe.3872

APPENDIX A: CALCULATION OF STATIC OPERATIONAL LOADS


Equivalent static wind loads were calculated according to the simplified procedure described by De Risi et al.14 The wind
load is proportional to the wind profile along the tower, 𝑣(𝑧), defined according to the relation proposed by ASCE/SEI
07-1052 :
( )0.2
𝑧
𝑣(𝑧) = 𝑣ℎ𝑢𝑏 ⋅ (A1)
ℎℎ𝑢𝑏

where ℎℎ𝑢𝑏 is the vertical distance between the hub centre and the water line, and 𝑣ℎ𝑢𝑏 is the wind velocity at this height.
The equivalent horizontal force, 𝐹(𝑧) can then be calculated:

𝐹(𝑧) = 0.5 ⋅ 𝜌𝑎𝑖𝑟 ⋅ 𝑣(𝑧)2 ⋅ 𝐴(𝑧) ⋅ 𝐶𝐷 (A2)


10969845, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/eqe.3872 by Cochrane Japan, Wiley Online Library on [03/04/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
24 HAMILTON and ABADIE

where 𝐴(𝑧) is the projected area of the tower section considered, and 𝐶𝐷 is the drag coefficient. Finally an additional con-
centrated wind load is applied at hub height. The force is evaluated through the formulation by Frohboese and Schmuck53
for wind speeds less than or equal to the rated wind speed, 𝑣𝑟 :

2
𝐹ℎ𝑢𝑏 = 0.5 ⋅ 𝜌𝑎𝑖𝑟 ⋅ 𝑣ℎ𝑢𝑏 ⋅ 𝐴𝑠𝑤𝑒𝑝𝑡 ⋅ 𝐶𝑇 (A3)

where 𝐴𝑠𝑤𝑒𝑝𝑡 is the area swept by the turbine blades and 𝐶𝑇 is calculated as:

1
𝐶𝑇 = 3.5 ⋅ (2 ⋅ 𝑣𝑟 + 3.5) (A4)
𝑣𝑟3

Equivalent static wave loads were calculated through integration of Morison’s equation:

1 𝜕𝑢
𝐹(𝑧) = 𝐶𝐷 𝜌𝐷|𝑢|𝑢 + 𝐶𝑀 𝜌𝐴 (A5)
2 𝜕𝑡

where 𝐹(𝑧) is the equivalent horizontal force at water depth 𝑧, 𝐷 is the monopile diameter, 𝐴 is its cross-sectional area,
𝐶𝐷 is the drag coefficient, and 𝐶𝑀 is the intertia coefficient. The water particle horizontal velocity, 𝑢 and acceleration are
undisturbed values that would apply at the member centreline, calculated according to linear Airy wave theory:
( )
𝜕𝜙 𝐻 𝑔𝑘 cosh(𝑘(𝑧 + 𝑑))
𝑢(𝑧) = − = 𝑆 cos 𝑘𝑥 − 𝜔𝑡 (A6)
𝜕𝑥 2𝜔 cosh 𝑘𝑑

where 𝜙 is the velocity potential, 𝐻𝑆 is the significant wave height, 𝑔 is the gravitational constant, 𝑘 = 2𝜋∕𝐿, 𝜔 is the wave
angular frequency, 𝑑 is the mean water depth, 𝑥 is the horizontal coordinate parallel to the direction of wave motion, and 𝑡
is time. Peak drag loading occurs as the wave passes the monopile, but the peak inertial load occurs at 𝑇∕4 before this time,
and contributes the vast majority of load. As such, the forces applied to the structure were calculated at time 𝑡 = 𝑇∕4. The
drag and inertia coefficients were determined according to the relationships between the Keulegan-Carpenter number,
𝐾𝑐 , steady-flow drag coefficient, 𝐶𝐷𝑆 , wake amplification factor, 𝜓 suggested in DNVGL-ST-0437.54 The monopile was
assumed to be a rough cylinder.

You might also like