You are on page 1of 13

Applied Ocean Research 33 (2011) 215–227

Contents lists available at ScienceDirect

Applied Ocean Research


journal homepage: www.elsevier.com/locate/apor

Incorporating irregular nonlinear waves in coupled simulation and reliability


studies of offshore wind turbines
Puneet Agarwal a,∗,1 , Lance Manuel b
a
Stress Engineering Services, Houston, TX, United States
b
Department of Civil, Architectural, and Environmental Engineering, University of Texas, Austin, TX, United States

article info abstract


Article history: Design of an offshore wind turbine requires estimation of loads on its rotor, tower and supporting
Received 13 June 2010 structure. These loads are obtained by time-domain simulations of the coupled aero-servo-hydro-elastic
Received in revised form model of the wind turbine. Accuracy of predicted loads depends on assumptions made in the simulation
29 November 2010
models employed, both for the turbine and for the input wind and wave conditions. Currently, waves
Accepted 2 February 2011
Available online 10 March 2011
are simulated using a linear irregular wave theory that is not appropriate for nonlinear waves, which
are even more pronounced in shallow water depths where wind farms are typically sited. The present
Keywords:
study investigates the use of irregular nonlinear (second-order) waves for estimating loads on the support
Offshore wind turbines structure (monopile) of an offshore wind turbine. We present the theory for the irregular nonlinear
Nonlinear irregular waves model and incorporate it in the commonly used wind turbine simulation software, FAST, which had been
Coupled simulation developed by National Renewable Energy Laboratory (NREL), but which had the modeling capability only
Long-term loads for irregular linear waves. We use an efficient algorithm for computation of nonlinear wave elevation
Statistical extrapolation and kinematics, so that a large number of time-domain simulations, which are required for prediction of
Inverse FORM long-term loads using statistical extrapolation, can easily be performed. To illustrate the influence of the
Convergence criteria alternative wave models, we compute loads at the base of the monopile of the NREL 5MW baseline wind
turbine model using linear and nonlinear irregular wave models. We show that for a given environmental
condition (i.e., the mean wind speed and the significant wave height), extreme loads are larger when
computed using the nonlinear wave model. We finally compute long-term loads, which are required for a
design load case according to the International Electrotechnical Commission guidelines, using the inverse
first-order reliability method. We discuss a convergence criteria that may be used to predict accurate
20-year loads and discuss wind versus wave dominance in the load prediction. We show that 20-year
long-term loads can be significantly higher when the nonlinear wave model is used.
© 2011 Elsevier Ltd. All rights reserved.

1. Introduction Response of an offshore wind turbine to the input wind and


waves, both of which are stochastic in nature, is estimated using
Offshore wind energy is becoming an important part of the coupled aero-servo-hydro-elastic simulation in time domain. Such
overall energy mix in Europe and has great potential within the time domain simulations are also used to obtain the data on load
United States and other parts of the world. The potential for extremes, which are needed for calculation of long-term extreme
offshore wind power in the United States, in water depths less than loads using the statistical extrapolation method as recommended
30 m where current fixed-bottom support structure technologies by the IEC 61400-3 standard [4] from the International Electrotech-
can be used, is about 98 GW [1]. Europe has a target of generating nical Commission (IEC). Statistical extrapolation refers to estima-
about 12% of its electricity from wind by the year 2020, of which tion of a rare load fractile associated with the desired long return
about one-third, or about 30 GW (gigawatts), is expected to come period (on the order of 20–50 years, typically). Several extrapola-
from offshore projects [2,3]. To meet such demand, it is important tion techniques, such as direct integration of short-term distribu-
to develop appropriate simulation tools for the design of offshore tions of load extremes (conditional on environmental conditions)
wind turbines. appropriately weighted by the likelihood of occurrence of the en-
vironmental conditions, as well as more efficient techniques such
as the inverse first-order reliability method (inverse FORM) have
∗ Corresponding author. Tel.: +1 2819552900; fax: +1 2819552638. been explored in wind turbine applications [5].
E-mail address: puneet.agarwal@stress.com (P. Agarwal). The accuracy of the turbine response, and of the long-term
1 Formerly, Graduate Research Assistant, University of Texas, United States. loads, depends on the proper modeling and interfacing of various
0141-1187/$ – see front matter © 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.apor.2011.02.001
216 P. Agarwal, L. Manuel / Applied Ocean Research 33 (2011) 215–227

parts of the physics (aeroelasticity, dynamics, control systems etc.) nonlinear waves. An important question regrading the accuracy
of an offshore wind turbine, as well as on the appropriate modeling of extrapolation is the required number of simulations for each
of the incident wind and waves. In this paper, we exclusively focus environmental state; we show that we use enough simulations
on how alternate wave models influence the turbine response. based on a convergence criterion developed recently [27]. Finally,
The current practice to model waves on offshore wind turbines we show how the nonlinear wave model affects turbine loads,
is limited to linear irregular waves, which are not an accurate particularly at larger wave heights, by focussing on occurrences of
representation of waves in shallow waters where offshore wind large loads in the response time series.
turbines are most commonly sited. Waves, in general, can be
nonlinear in nature, and nonlinearity is even more pronounced 2. Hydrodynamics in coupled simulation of offshore wind
in shallow water depths. Therefore we attempt to incorporate a turbines
nonlinear irregular wave model in the coupled aero-servo-hydro-
elastic simulation of offshore wind turbines.
Structural response of offshore wind turbines in the time do-
Among several nonlinear irregular wave models available in
main is analyzed using coupled aero-servo-hydro-elastic simula-
the literature, the one that has been recommended by offshore
tion programs. One such program is FAST (Fatigue, Aerodynamics,
guidelines [6] and that has been increasingly applied to a variety
Structures and Turbulence) [25], developed at the National Renew-
of problems in recent years [7–9,6] is the second-order nonlinear
able Energy Laboratory (NREL), which models a wind turbine as
irregular wave model developed by Sharma and Dean [10] for
a combination of rigid and flexible bodies wherein flexible beam
finite water depths. This second-order nonlinear irregular wave
elements (representing the blades and the tower) are formulated
model is based on the solution of the Laplace equation in
in terms of generalized coordinates. Aerodynamic forces on the
velocity potential, associated with nonlinear boundary conditions,
rotating blades of a horizontal-axis wind turbine are commonly
using a second-order perturbation expansion of the relevant
computed using the blade element momentum theory [28]. FAST
variables (velocity potential and sea surface elevation) and a
models various control systems including blade and pitch control,
Taylor series expansion of nonlinear boundary conditions about
yaw control, and high-speed shaft brake control. FAST uses an ir-
the free surface. Such an approach was first presented by Longuet-
regular linear wave model with Morison’s equation to compute
Higgins [11] and Hasselmann [12] for infinite water depths. Since
hydrodynamic loads for offshore wind turbines. Perhaps the most
its introduction in 1979, the second-order nonlinear irregular wave
important aspect of the FAST program is the proper interfacing of
model developed by Sharma and Dean has been studied by several
various parts of the physics—aeroelasticity, dynamics, control sys-
researchers. Hu and Zhao [13], Langley [14], Longuet-Higgins [15]
tems, hydrodynamics, etc.—of an offshore wind turbine. Without
and Forristall [16], among others, investigated the statistical
such coupling, it is not possible to accurately assess the response
properties (skewness in particular) of this nonlinear irregular wave
of wind turbines. We next discuss hydrodynamic calculations in
model. Marthinsen and Winterstein [17] and Stansberg [18] also
FAST, while highlighting its limitations to model shallow water
studied statistical properties of a very similar second-order wave
waves.
model. Forristall [19] compared various stretching techniques
for calculating wave kinematics above the mean sea level. Due
to the wide use and adoption of this nonlinear irregular wave 2.1. Irregular linear wave simulation
model and because of its theoretical basis (as distinct from some
semi-empirical models, such as the New Wave model [20], the In the current version of the FAST program, waves are modeled
constrained New Wave model [21] and a hybrid wave model [22]), using the irregular linear wave model, which is the most common
and also due to its computational efficiency for simulations model to represent stochastic (or irregular) ocean waves. A flow-
(as distinct from more realistic though computationally very chart for calculation of wave kinematics using the irregular
expensive models based on Boussinesq theory [23,24]), we will use linear wave model for unidirectional waves and the resulting
this second-order nonlinear irregular wave model, as developed hydrodynamic loads on the monopile support structure of the
by Sharma and Dean [10], to simulate nonlinear irregular waves in offshore wind turbine is shown in Fig. 1. These calculations are
shallow water depths for offshore wind turbine loads predictions. rather standard and may be found in several well-established
The outline of this paper is as follows: we first present an references [29,30], so we discuss them only briefly here.
overview of the hydrodynamics—employing the currently used The starting point in simulation of irregular waves is a (one-
linear irregular wave model—in the coupled simulation of offshore sided) wave spectrum, S (ωm ), where ωm refers to the frequency of
wind turbines. We present the theoretical formulation of the the mth wave component; the integer, m, refers to a frequency in-
second-order nonlinear irregular wave model. We limit our focus dex that ranges from 1 to N. Random seeds are used to generate the
to unidirectional (long-crested) waves. We then present a detailed random phase, φm , which are assumed to be uniformly distributed
procedure, which we have incorporated in the turbine response over [0, 2π ], and Rayleigh distributed amplitudes, Am , whose mean
simulation software, FAST [25], for efficient simulation of the sea square value, E [A2m ], is 2S (ωm )∆ω. Fourier coefficients, X (ωm ), for
surface elevation and the water particle kinematics process based the sea surface elevation process, η1 (x, tp ), (the subscript ‘‘1’’ de-
on the nonlinear irregular wave model. We use a utility-scale notes the linear, or first-order, wave model) are then computed.
5 MW offshore wind turbine model developed at the National Here, the spatial location is denoted by x (see Fig. 2 for a schematic),
Renewable Energy Laboratory (NREL) [26] to compare turbine and the time-instant is denoted by tp = p∆t, where ∆t = T /N,
response (mainly the fore-aft tower bending moment at the such that p = 1, 2, . . . , N. Fourier coefficients for the water par-
mudline) due to linear and nonlinear irregular waves. We compute ticle velocity, Xu (h, z , ωm ), and for the water particle acceleration,
the turbine response for a representative environmental state, Xu̇ (h, z , ωm ), are then computed using formulae from the irregular
and show that when nonlinear waves are used, turbine loads linear wave model as shown in the flow-chart. Once these are avail-
can be larger and tower dynamics can be influenced to some able, time-series of sea surface elevation, and water particle veloc-
extent as well. We briefly discuss the procedure for extrapolated ity and acceleration, are computed using the Inverse Fast Fourier
long-term load prediction using the inverse FORM technique. We Transform (IFFT) algorithm. Hydrodynamic forces per unit length,
then compare the long-term loads based on linear and irregular f (0, z , t ) (for simplicity of notation, we will drop the argument
wave models. We discuss short-distributions of turbine loads (0, z , t ) hereon), at any node located at a depth z, along the center-
for governing environmental states, and highlight the effect of line of the monopile, which we arbitrarily chose at x = 0 (Fig. 2)
P. Agarwal, L. Manuel / Applied Ocean Research 33 (2011) 215–227 217

Fig. 2. Schematic illustrating the spatial coordinate, x, the vertical coordinate, z,


and the sea surface elevation, η(x, t ), relative to the mean sea level (MSL).

3.1. Theoretical model

The nonlinear sea surface elevation, η(t ), may be expressed as


a sum of first- and second-order components, such that η(t ) =
η1 (t ) + η2 (t ). The first-order component, η1 (t ), is expressed as in
Fig. 1. Flow-chart for calculation of irregular linear wave kinematics and linear wave theory by
hydrodynamic loads on the monopile support structure in FAST.
N

are computed using the Morison’s equation [29] as follows: η1 (t ) = Am cos(ωm t − φm ) (2)
1 m=1
f = fD + fM = CD ρ D(u − q̇)|u − q̇|
2 where ωm refers to the frequency of the mth wave component. The
random phase, φm , and the amplitudes of the wave components,
πD 2
πD 2
[ ]
+ CM ρ u̇ − (CM − 1) ρ q̈ (1) Am , have been defined in Fig. 1.
4 4 The second-order component, η2 (t ), is obtained as a result of
where fD and fM are drag and inertia forces, respectively. Also, CD the interactions of sums and differences of frequencies as follows:
and CM are the drag and the inertia coefficients, respectively; ρ is N −
− N
the density of water, and D is the diameter of the cylinder. The vari- η2 (t ) = mn cos(ψm − ψn ) + Bmn cos(ψm + ψn )}
Am An {B− +
 
(3)
ables u(= u1 ) and u̇(= u˙1 ) are the undisturbed water particle ve- m=1 n=1
locity and acceleration, respectively. Variables q̇ and q̈ denote the
velocity and acceleration of the corresponding node on the struc- where ψm = (ωm t − φm ) and the second-order transfer functions,
ture, which are obtained from the dynamic analysis of the entire B− +
mn and Bmn , are obtained from the solution of Laplace’s equation
turbine system, including blade aero-elasticity, structural dynam- for the velocity potential with nonlinear boundary conditions. They
ics, control actions, generator, etc. at each time step. Note that such (i.e., B− +
mn and Bmn ) are functions of frequency and wave number and
aero-servo-elastic analysis needs to be ‘‘coupled’’ with the hydro- are independent of the spectrum used.
dynamic analysis in order to correctly account for the relative mo- The velocity potential, 8, is comprised of first and second-order
components such that 8 = 81 + 82 . These first and second-order
tion between water particles and the monopile in calculation of
velocity potentials are given as follows:
hydrodynamic loads according to Eq. (1).
One limitation of the existing FAST program is the use of the N
− cosh(km (h + z ))
linear irregular wave model, which is not appropriate for shallow 81 = bm sin ψm (4)
cosh(km h)
water depths, where waves are irregular as well as nonlinear. m=1

We address this model limitation by incorporating a second- cosh(k±


mn (h + z ))
N N

[ ]
1 −−
order nonlinear irregular wave model in the FAST program for 82 = bm bn mn
sin(ψm ± ψn ) (5)
4 m=1 n=1 cosh(k±
mn h) (ωm ± ωn )
the coupled aero-servo-hydro-elastic simulation of offshore wind
turbines. where bm = Am g /ωm and k± mn = |km ± kn |. Also, the linear dis-

3. Second-order nonlinear irregular waves persion relation, ωm 2


= gkm tanh(km h), relates the wave number,
km , to the frequency, ωm , where h is the water depth and g is accel-
Linear wave theory for regular or irregular waves involves the eration due to gravity. Expressions for the transfer functions, B± mn
solution of Laplace’s equation expressed in terms of a velocity and D± mn , appearing in Eqs. (3) and (5), respectively, were derived
potential and the use of linearized boundary conditions [29]. For by Sharma and Dean[10], and are also summarized in Appendix A.
nonlinear waves, the theory involves application of a perturbation The horizontal water particle velocity, u(z , t ), and the horizontal
approach to solve Laplace’s equation with nonlinear boundary water particle acceleration u̇(z , t ) may be obtained from the ve-
conditions. Sharma and Dean [10] used such an approach to derive locity potential by taking derivatives such that
a nonlinear wave theory for finite water depths. We will use the
u(z , t ) = ∂ 8/∂ x, u̇(z , t ) = ∂ u(z , t )/∂ t . (6)
formulation of Sharma and Dean, which is described very briefly
below. This theory is also recommended in some guidelines for Second-order waves are thus obtained as a result of sum and
offshore structures [6,9]. difference interactions between pairs of frequencies. The phases
218 P. Agarwal, L. Manuel / Applied Ocean Research 33 (2011) 215–227

of the second-order contributions are also determined by sum and similar to that for linear waves, can be used to perform the
difference interactions of the phases of the first-order component nonlinear wave simulations more efficiently.
phases, which are random. The double sum in Eq. (3), to simulate the second-order
component of the sea surface elevation process, may be rewritten
3.1.1. Statistics of nonlinear sea surface elevation as
 
Langley [14] developed a procedure, based on Volterra series N −
− N
η2 (tp ) = ℜ ±
exp(−i(m ± n)∆ωp∆t )
 
models, to compute the statistical moments of a nonlinear sea Xmn (7)
surface elevation process from wave spectrum and associated m=1 n=1
second-order transfer functions. The variance of the nonlinear
mn exp(−i(φm ± φn ))
±
Xmn = Am An B± (8)
wave process is sum of the variances of the first-order (linear) and
second-order processes. The variance of the first-order process, η1 , where tp = p∆t, ωm = m∆ω and ωn = n∆ω; m, n, p = 1,
is still the area under the power spectral density function of the 2, . . . , N. Also, ∆t = T /N, where T is the period (duration) of
sea surface elevation process, S (ωm ), which can be a function of the simulations and ∆ω = 2π /T . In simulations of waves, an
the significant wave height as is the case with parametric spectra appropriate cut-off frequency, ωcut , must be used. We have used
such as JONSWAP. The variance of the second-order process, η2 , is ωcut = ns ωP , where ns is factor used to the scale the peak spectral
comparatively smaller generally. Skewness of the nonlinear wave frequency, ωP , in order to find the cut-off frequency. Based on
process is entirely due to the second-order component, and, in fact, the suggestion in [35], ns is chosen as 3.0 here for both linear
nonlinearity described by the second-order irregular wave model and nonlinear waves. Using an appropriate cut-off frequency is
is directly related to the process skewness. One can compute the especially important for simulation of nonlinear waves, because
variance and skewness from wave spectra that should match that if the cut-off frequency is too high, sum-frequency interactions
from a simulated nonlinear sea surface elevation ensemble. from high-frequency first-order components can lead to events
with too high steepness. This was discussed by Stansberg [36], who
3.1.2. Validity of the second-order wave model also proposed another criterion to choose the cut-off frequency for
While the second-order irregular wave model is a more nonlinear waves.
accurate representation of irregular seas in shallow waters Performing this sum (as given by Eq. (7)) is computationally
compared to the linear irregular wave model, it obviously does not inefficient even for moderate values of N, as a total of N 2 terms
model the complete nonlinear character of waves, and is not valid are to be summed. Therefore, we seek an implementation based
for all cases. The physical parameter that determines the range of on the IFFT (inverse Fast Fourier Transform) technique. We start by
validity is the wave steepness, with some consideration for water rewriting the double summation in Eq. (7) as an equivalent single
depth as well, especially in shallow water depths. When the wave summation as follows:
steepness exceeds a certain value, the second-order model is no  
2π j
N
− 
longer valid, and a higher-order model is required. In fact, when η2 (tp ) = ℜ ±
Yj exp −i p . (9)
waves become too steep, they can break and no model based on j =1
N
solution of Laplace’s equation (in terms of velocity potential) is
valid. The wave steepness, s, according to DNV guidelines [6], is Note that the term within the summation in Eq. (9) is the definition
defined as s = Hs /Lz , where Hs is the significant wave height of the inverse discrete Fourier transform of Yj± , which can be
and Lz is the wavelength based on the mean zero-crossing period efficiently computed using Inverse Fast Fourier transform (IFFT)
(obtained using the linear dispersion relation). techniques, such that
Based on results from numerical simulations, Hu and Zhao [13] η2 (tp ) = ℜ[IFFT (Yj± )]. (10)
suggested that the second-order wave model presented above is
valid as long as the wave steepness is smaller than approximately ±
The Fourier coefficients Yj are obtained by ‘‘equating’’ Eq. (9) to
0.08. This empirical limit can be used to ascertain whether the Eq. (7). This involves conversion of a two-dimensional sum over
second-order wave model should be used for a given significant N 2 terms to a one-dimensional sum over N terms by collecting all
wave height and wavelength combination. (m, n) pairs that yield a sum, or difference, equal to j, for sum-
and difference-frequency interactions, respectively. For example,
3.2. Numerical implementation (m, n) = (1, 3), (2, 2) and (3, 1) would all yield a sum for j = 4; so,
these three Xmn+
contributions must be collected to form Y4+ . Note
While the simulation of second-order wave model has been that both the time index, p, and the frequency indices, j, still range
discussed in several published studies [31–33,18,16,6], a complete from 1 to N.
recipe for the numerical simulation of such second-order waves For sum-frequency interaction, as m and n range from 1 to M,
is not yet available in one place. This is in contrast to the case for the sum, (m + n), ranges from 2 to 2M. Note that more than
linear irregular waves, for which a recipe for numerical simulation one (m, n) pair may result in the sum, j = m + n. The Fourier
with all the pertinent details is available in several places [29,34]. coefficients Yj+ , for 2 ≤ j ≤ 2M are obtained as
Therefore, we discuss the numerical simulation of second-order −−
waves in detail below. Yj+ = +
Xmn (11)
A numerical simulation of irregular (random) linear or first-
  
m+n=j
order waves, which involves a single summation (Eq. (2)),
can be efficiently performed using the Inverse Fast Fourier with Y1+ = 0.
Transform (IFFT). On the other hand, simulation of random For difference-frequency interactions, as m and n range from 1
nonlinear or second-order waves according to Eq. (3) involves a to M, the difference, (m − n), ranges from −(M − 1) to (M − 1). Note
double summation, which can be very expensive. However, one that Eq. (7) requires only cosine terms to be computed, and since
can rewrite the double summation as a single summation by cos(ψm −ψn ) = cos(−(ψm −ψn )), negative difference-frequencies
appropriately re-assembling and rewriting indices (or coefficients) have the same effect as positive difference-frequencies. Therefore,
in the double summation. Once the indices for an equivalent single we need to concern ourselves only with cos(|ψm − ψn |) or,
summation are assembled, a one-dimensional IFFT procedure, effectively, with |m − n|. When creating the terms Yj− , we need
P. Agarwal, L. Manuel / Applied Ocean Research 33 (2011) 215–227 219

to keep the consistent phases, φmn



(= φm − φn ), to make sure Table 1
the IFFT-based simulation (Eq. (9)) is the same as the summation- Comparison of statistics of the sea surface elevation process simulated using linear
and nonlinear irregular wave models for a JONSWAP spectrum with Hs = 7.5 m
based simulation (Eq. (7)). Following from the symmetry of the and Tp = 12.3 s.
cosine function, we need
Sea-surface elevation statistics Wave model
(φm − φn ), m>n

Linear Nonlinear
φmn

= (12)
−(φm − φn ), m ≤ n. Std. dev. (m) 1.8 2.0
Max (m) 5.6 6.6
The Fourier coefficients Yj for 1 ≤ j ≤ (M − 1) are obtained as

Skewness 0.0 0.1
Kurtosis 2.9 3.2
Peak factora
−−
3.0 3.3
Yj− = −
Xmn (13)
a
   10-min median extreme peak factor (computed based on 50 simulations).
|m−n|=j

with Yj− = 0 for j ≥ M. Table 2


Using the procedure described above, we can obtain the Comparison of maximum water particle velocity, umax , acceleration, u̇max , at the
free surface, averaged over fifty simulations, computed with linear and nonlinear
coefficients, Yj± , for Eq. (9) from Eqs. (11) and (13), and simulate
irregular waves using a JONSWAP spectrum with Hs = 7.5 m and Tp = 12.3 s.
the time series of the second-order sea surface elevation, η2 (tp ),
Depth = 20 m Depth = 30 m
using Eq. (10).
umax u̇max umax u̇max
Second-order components of the wave kinematics, such as the
(m/s) (m/s2 ) (m/s) (m/s2 )
horizontal particle velocity at x = 0 and at a depth z, u(0, z , tp ), and
the horizontal particle acceleration, u̇(0, z , tp ), can be simulated by Linear waves 5.05 3.64 4.59 3.81
Nonlinear waves 7.05 5.20 5.38 4.10
a similar approach. Again, to simplify notation, we denote these
Ratio 1.40 1.43 1.17 1.08
variables as u(z , tp ) and u̇(z , tp ). By substituting the second-order
velocity potential (Eq. (5)) in the definition of particle kinematics
given by (Eq. (6)), we obtain the following expressions for velocity only additional input to be provided by the user. Because of the use
and acceleration: of the efficient simulation algorithm described above, the compu-
 ] tation time for nonlinear irregular model is similar to that for the

N −
N [ 
− linear irregular waves, and a large number of time-domain simu-
u2 (z , tp ) = ℜ Umn exp −i(m ± n)
±
p (14)
N lations required for load extrapolation can easily be performed.
m=1 n=1
 ]

N −
N [
3.3. Wave simulation example
− 
u̇2 (z , tp ) = ℑ U̇mn exp −i(m ± n)
±
p (15)
m=1 n=1
N
We simulate linear and nonlinear waves with a significant wave
where ℜ(·) and ℑ(·) denote the real and imaginary components of height, Hs , of 7.5 m. We assume a peak spectral period, Tp , of
the argument, respectively, and 12.3 s, and use a JONSWAP spectrum. A water depth of 20 m
is assumed (same as for the wind turbine model discussed later
±
Umn ±
= Zmn exp(−i(φm ± φn )sgn(m ± n)) (16) in the paper). Based on 50 simulations, each of 10-min dura-
tion, the average value of statistical moments, maxima and ex-
U̇mn = −(ωm ± ωn )Umn
± ±
(17)
treme peak factors for the sea surface elevation process, simulated
1 cosh(kmn (h + z ))
± ±
Dmn using linear and nonlinear irregular waves, are summarized in
mn .
±
Zmn = bm bn k± (18) Table 1. Statistical moments are computed from the simulated time
4 cosh(kmn h)
±
(ωm ± ωn )
series, and they were found to match the target moments com-
Also, sgn(q) denotes the signum function such that puted from a theoretical formulation presented by Langley [14].
Nonlinear waves have a non-zero skewness and a kurtosis larger
−1 if q < 0

than three, which indicates the non-Gaussian character of these
sgn(q) = 0 if q = 0 (19)
nonlinear waves. Because of the larger skewness and kurtosis
1 if q > 0.
and, hence the somewhat larger peak factor (= [Max − Mean]/
The simulation of particle velocity and acceleration according to [Standard Deviation]), associated with nonlinear waves, extremes
Eqs. (14) and (15), respectively, requires a double summation. We based on the use of a nonlinear wave theory are larger than those
can reduce these double summations to single summations just as based on linear waves. Since nonlinear waves tend to have sharper
we did for sea surface elevation, and can readily simulate particle crests, the maximum of the sea surface elevation is larger (by about
velocity and acceleration according to the following equations: 1 m or 18%) than is the case for linear waves.
Estimates of the maximum horizontal particle velocity, umax ,
u2 (z , tp ) = ℜ[IFFT (Wj± )] (20) and acceleration, u̇max , at the free surface are presented in Table 2.
It is clearly seen that maximum velocity and acceleration values
u̇2 (z , tp ) = ℑ[IFFT (Ẇj± )]. (21) increase when nonlinear waves are modeled, and the increase is
± ± about 40% for the water depth of 20 m. Table 2 also shows that
The Fourier coefficients, Wj , in Eq. (20) and Ẇj in Eq. (21), are
± ±
influence of nonlinear waves decreases with increase in water
assembled from the coefficients, Umn and U̇mn , respectively, using depth. This is because at larger water depths, waves tend to
exactly the same approach as for the coefficients, Yj± , in Eq. (9) that become more and more linear.
±
were in turn assembled from the coefficients, Xmn , in Eq. (7). As a result of larger particle velocities and accelerations, the
A flowchart for calculation of irregular nonlinear wave kine- hydrodynamic loads, and hence the lateral base forces, on a
matics and resulting hydrodynamic loads is described in Fig. 3. monopile would be larger for nonlinear waves, compared to those
These calculations are implemented in the computer program FAST for linear waves. Furthermore, due to larger wave heights for
for coupled aero-servo-hydro-elastic analysis of offshore wind tur- nonlinear waves, it is expected that a greater portion of the
bines. A switch for irregular wave model (linear or nonlinear) is the monopile would get submerged and, as a result, the loads would be
220 P. Agarwal, L. Manuel / Applied Ocean Research 33 (2011) 215–227

Fig. 3. Flow-chart for calculation of second-order nonlinear irregular wave kinematics and hydrodynamic loads on the monopile support structure in FAST.

further amplified. For a more detailed discussion of the influence manufactured today is considered here. The turbine is a variable-
of nonlinear waves on drag and inertia forces, and how forces vary speed, collective pitch-controlled machine with a maximum rotor
for a range of monopile diameter and water depths, the reader is speed of 12.1 rpm; its rated wind speed is 11.5 m/s. It is assumed
referred to a previous study by authors [37]. to have a hub height of 90 m above the mean sea level, and a
rotor diameter of 126 m. It is assumed to be sited in 20 m of
4. Effect of wave model on turbine response water; it has a monopile support structure of 6 m diameter, which
is assumed to be rigidly connected to the seafloor. The turbine is
We now investigate how the alternative wave models discussed assumed to be installed at an IEC Class I-B wind regime site [4]. A
above affect the response of an offshore wind turbine. We focus Kaimal power spectrum and an exponential coherence spectrum
on the fore-aft tower bending moment and the shear force at the are employed to describe the inflow turbulence random field over
mudline, as these are directly affected by waves. Loads on the the rotor plane, which is simulated using the computer program,
turbine rotor, on the other hand, are not affected as much by waves TurbSim [38].
as by wind. In the following, we first discuss the turbine model For the hydrodynamic loading on the support structure, ir-
used, and then compare time-series and statistics of the fore-aft regular long-crested waves are simulated using a JONSWAP
tower bending moment resulting from linear and nonlinear wave spectrum [39]. This same wave spectrum is used for sim-
models. We finally compute loads for a return period of 20 years in ulating linear and nonlinear irregular waves. Hydrodynamic
order to assess influence of the alternative wave models on long- loads are computed using Morison’s equation (Eq. (1)). Wheeler
term loads. stretching [29], whose use has been suggested for second-order
waves [6,40] as well, is used to represent water particle kinemat-
4.1. Turbine model
ics and hydrodynamic loads up to the changing instantaneous sea
A 5 MW wind turbine model developed at NREL [26] surface. Note that other techniques [19,36] for stretching of wave
closely representing utility-scale offshore wind turbines being kinematics are also available, and these could also be used.
P. Agarwal, L. Manuel / Applied Ocean Research 33 (2011) 215–227 221

Table 3 twice the spectral peak frequency of 0.08 Hz (since Tp = 12.3 s),
Comparison of 10-min statistics, averaged over 100 simulations, of the fore-aft and another small peak close to a zero frequency. Such secondary
tower base shear (FATBS) and the fore-aft tower bending moment (FATBM) at the
mudline for linear and nonlinear waves, for a JONSWAP spectrum with Hs = 7.5 m
peaks arise due to sum and difference interactions of frequencies
and Tp = 12.3 s, and wind speed, V = 18 m/s. according to the second-order nonlinear wave model (see Eq. (3)).
These secondary peaks also appear in power spectra of both the
FATBS (MN) FATBM (MN m)
tower base shear and the tower bending moment (Fig. 4b). The
Wave model → Linear Nonlinear Linear Nonlinear
tower bending moment power spectrum has a significant peak at
Max 3.06 3.88 90.43 98.58 around 0.27 Hz, which is the natural frequency for the first bending
Std. dev. 0.81 0.86 13.43 14.04
mode of vibration of the tower in the fore-aft direction; the peak is
Skewness 0.08 0.25 0.15 0.22
Kurtosis 3.04 3.81 3.23 3.60 somewhat suppressed but that is due to the aerodynamic damping
Peak factor 3.32 4.08 3.70 4.10 from turbine rotor [41–43]. These observations suggest that tower
dynamics, which may be important in the overall response, may be
influenced by the nonlinear waves.
4.2. Turbine response Using the statistics, time-series and power spectral density of
load process, we have shown that the nonlinear waves, which
We consider environmental conditions involving a significant better represent waves in the shallow water depths, can influence
wave height of 7.5 m with a mean wind speed of 18 m/s. This the tower loads. We next discuss how nonlinear waves affect long-
combination of V and Hs governs long-term fore-aft tower bending term tower loads, which are relevant in the design of offshore wind
moment at the mudline for a return period of 20 years; we will turbines.
discuss long-term loads later in the paper, and for now, we discuss
how linear and nonlinear irregular waves influence the turbine
5. Long-term loads
response.
Table 3 shows statistics of the fore-aft tower base shear (FATBS)
and the fore-aft tower bending moment (FATBM) at the mudline, In the design of offshore (as well as onshore) wind turbines,
averaged over 100 10-min simulations. The 10-min maximum one is required to compute long-term loads for a return period
FATBM with the nonlinear waves is 99 MN m, which is about 9% in the order of 20 or 50 years. Such long-term loads, according
larger than that due to the linear waves. The maximum FATBS due to the design standard [4] from the International Electrotechnical
to the nonlinear waves is, however, almost 27% larger than that due Commission (IEC), are to be computed using the statistical
to the linear waves. Note that because of a larger lever arm (hub extrapolation method. The load extremes data required for
height of 90 m), aerodynamic forces on the turbine rotor have a the statistical extrapolation are obtained by stochastic turbine
greater influence on the fore-aft tower bending moment than they response simulations in time domain. While we have so far
have on the fore-aft tower base shear (see [41] for more details on discussed how nonlinear waves influence the turbine response for
how wave nonlinearity affects tower loads in presence and absence a given environmental state, we now discuss how long-term loads
of wind). Shear forces at the base of the tower are influenced differ when irregular waves are modeled using linear or nonlinear
more directly by hydrodynamic loads, hence the larger sea surface theory.
elevation and particle kinematics due to nonlinear waves result in In the statistical extrapolation using the so-called direct
a greater increase in base shear than in bending moment. integration method, one estimates the turbine nominal load for
We can further understand hydrodynamic loads by studying design, lT , associated with an acceptable probability of exceedance,
the non-Gaussian character in these load processes. The degree to PT , or, equivalently, with a target return period of T years, as
which a process is non-Gaussian relates to the extent by which its follows:
skewness deviates from zero and its kurtosis deviates from three.

A process with a positive skewness and a kurtosis greater than PT = P [L > lT ] = P [L > lT |X = x]fX (x)dx (22)
X
three (as is the case here) will generally lead to larger peak factors.
This will, in turn, result in larger extremes associated with any where fX (x) represents the joint probability density function of
specified rare probability level compared to those predicted for a the environmental random variables, X , and L represents the load
Gaussian process. Note that it is such low exceedance probability measure of interest. For different trial values of the load, lT , Eq. (22)
levels that are of eventual interest in predicting long-term loads enables one to compute the long-term probability by integrating
for ultimate limit states. Table 3 shows that skewness, kurtosis the short-term load exceedance probability conditional on X ,
and peak factor estimates from the simulations are always larger i.e., P [L > lT |X = x], with the relative likelihood of different values
with nonlinear waves. As a result, extreme loads predicted based of X . This method, while exact, is expensive as one is required
on the use of nonlinear waves will also be larger. Furthermore, to integrate over the entire domain of all the environmental
increments in skewness, kurtosis and peak factor (or deviations random variables. In this study, two environmental random
from the Gaussian) due to nonlinear, and non-Gaussian, waves are variables comprise X ; these are the 10-min average wind speed,
larger for the base shear than for the bending moment; this, again, V , at hub height in the along-wind direction and the significant
is because waves affect base shear more directly. wave height, Hs , for waves assumed to be aligned with the
We study a representative 200-s segment from a single 10- wind.
min simulation time history to further understand the influence The direct integration method is numerically very expensive.
of wave nonlinearity. Fig. 4a shows that crests of the sea surface An alternative to compute long-term loads is the inverse first-order
elevation process, and those of fore-aft tower bending moment reliability method (FORM) [44], which has been shown in an earlier
and base shear are systematically higher for the nonlinear wave study [45] to be as accurate as the direct integration method for the
model than for the linear model. Moreover, the maxima of tower same offshore wind turbine. In the inverse FORM, for the present
loads occur almost at same instant as the maxima of sea surface application, one considers a surface in a three-dimensional space,
elevation, which clearly indicates that the maxima of tower loads Y = (V , Hs , L), of physical random variables, on one side of which
are influenced by waves. (i.e., the ‘‘failure’’ side), it is assumed that L > lT . It is possible to
The power spectrum of the sea surface elevation (Fig. 4b) for the mathematically transform this space to an independent standard
nonlinear case shows a secondary peak at about 0.16 Hz, which is normal space U = (U1 , U2 , U3 ). A sphere of radius, β , in the
222 P. Agarwal, L. Manuel / Applied Ocean Research 33 (2011) 215–227

(a) A 200-s segment of time series.

(b) Power spectral densities (PSD).

Fig. 4. (a) Time series and (b) power spectral densities of the wave elevation, wind speed, fore-aft tower base shear (FATBS) and fore-aft tower bending moment (FATBM)
at the mudline for V = 18 m/s, Hs = 7.5 m (Tp = 12.3 s).

standard normal space is defined as follows: If at any point on this sphere, a tangent plane were drawn, the
probability of occurrence of all points on the side of this plane away
u21 + u22 + u23 =β .
2
(23)
P. Agarwal, L. Manuel / Applied Ocean Research 33 (2011) 215–227 223

(a) V = 16 m/s, Hs = 5.5 m. (b) V = 18 m/s, Hs = 7.5 m.

Fig. 5. Empirical probability distributions of 10-min maxima of the fore-aft tower bending moment at the mudline based on 500 10-min simulations.

(a) V = 16 m/s, Hs = 5.5 m. (b) V = 18 m/s, Hs = 7.5 m.

Fig. 6. Empirical probability distributions of 10-min maxima of the wave height based on 500 10-min simulations for (a) V = 16 m/s, Hs = 5.5 m; and (b) V = 18 m/s,
Hs = 7.5 m.

from the origin is 8(−β). It is noted here that β is directly related Before presenting the actual long-term 20-year loads, we first
to the target probability of load exceedance; namely, PT = 8(−β), compare short-term distributions for these two environmental
where 8() represents the cumulative distribution function of a states in the following.
standard normal random variable. Since each point on the sphere
is associated with the same reliability index, the desired load, lT , 5.1. Short-term distributions
is also associated with this same reliability. The transformation
of the random variables involved from the physical space, Y , to Fig. 5 shows empirical (short-term) probability distributions
the standard normal U space is carried out via the Rosenblatt of 10-min load (fore-aft tower bending moment at the mudline)
transformation such that FV (v) = 8(u1 ), FH |V (h) = 8(u2 ), and maxima when linear and nonlinear waves are used for (a) the
FL|V ,H (l) = 8(u3 ), where F () denotes the cumulative distribution governing environmental state for the linear wave model (V =
function in each case. A point on the sphere defined by Eq. (23) 16 m/s, Hs = 5.5 m), and (b) the governing environmental state
where the load attains its maximum value is the ‘‘design’’ point, for the nonlinear wave model (V = 18 m/s, Hs = 7.5 m). The
and this load represents the desired nominal T -year return period distributions of load maxima for linear and nonlinear waves are
load, lT . quite different for V = 18 m/s and Hs = 7.5 m while they are
According to inverse FORM, we estimate the load for the fractile almost identical for V = 16 m/s and Hs = 5.5 m. It is interesting
level, p3 = 8(u3 ), for all possible (V , Hs ) pairs, and the largest of to note that these trends are similar for short-term distributions of
such load fractiles is the nominal load, lT . Based on Eq. (23), the 10-min maximum wave heights (crest to trough), which are shown
fractile level, p3 , corresponding to the target reliability index, β , is in Fig. 6. This is because hydrodynamic loads on this monopile
obtained as follows: (of 6 m diameter) are governed by inertia forces (see [37] for
 details), which are directly proportional to wave heights. It is worth
p3 = Φ ( β 2 − [8−1 (FV (v))]2 − [8−1 (FH |V (h))]2 ). (24) mentioning that the differences resulting from using linear and
nonlinear wave models are not as pronounced for loads as for wave
Calculations based on the inverse FORM show that a mean heights, because loads on this wind turbine are primarily driven
wind speed, V , of 18 m/s and a significant wave height, Hs , of 7.5 by wind and waves are only secondary in importance. Note that
m govern the long-term 20-year load when the nonlinear wave we have used 500 simulations here since the desired exceedance
model is used. This is different from the governing values for V of probability (= 1 − p3 ) for V = 18 m/s and Hs = 7.5 m is 2.03 ×
16 m/s and for Hs of 5.5 m when a linear wave model was used; 10−3 , which could be empirically observed, even if only once, with
which is discussed in detail in a previous study by authors [45]. 500 simulations. For comparison, we also ran 500 simulations for
224 P. Agarwal, L. Manuel / Applied Ocean Research 33 (2011) 215–227

Table 4 Table 5
Estimates of the 84th percentile 10-min maximum fore-aft tower bending moment Estimates of the 90% relative confidence bound (RCB) on the 84th percentile 10-min
at the mudline from 10 separate sets of 50 simulations each. maximum fore-aft tower bending moment at the mudline.
84th percentile 10-min maximum fore-aft tower 90% RCB on 84th percentile 10-min maximum
bending moment from 50 simulations (in MN m) fore-aft tower bending moment (in %)
V = 16 m/s, Hs = 5.5 m V = 18 m/s, Hs = 7.5 m V = 16 m/s, Hs = 5.5 m V = 18 m/s, Hs = 7.5 m
Waves: Linear Nonlinear Linear Nonlinear Waves: Linear Nonlinear Linear Nonlinear
Set no. # sims.

1 100.1 100.0 97.6 108.2 50 6.94 6.12 7.77 8.55


2 98.6 99.6 97.8 108.4 500 2.30 1.84 2.59 3.07
3 101.2 101.5 98.8 113.3
4 97.8 100.2 98.4 106.0
5 104.8 107.3 96.4 110.2 Table 6
6 98.8 98.8 93.2 107.0 Extrapolated 20-year loads (10-min maximum fore-aft tower bending moment at
7 100.0 100.2 94.4 106.4 mudline) from 50 simulations.
8 98.6 100.4 96.1 108.2
Extrapolated 10-min maximum fore-aft tower
9 99.5 100.5 94.9 106.8
bending moment (in MN m)
10 98.9 100.2 96.7 111.8
V = 16 m/s, Hs = 5.5 m V = 18 m/s, Hs = 7.5 m
Average 99.8 101.0 96.4 108.6
Lin. Nonlin. Ratio Lin. Nonlin. Ratio
Range (%) 7.0 8.4 5.7 6.7
Min 116.9 118.5 1.01 110.8 125.0 1.13
Note: Range = 100 × (Max − Min)/Median, over 10 sets.
Max 132.9 137.8 1.04 121.2 142.4 1.17
Median 125.6 128.1 1.02 117.7 138.4 1.18
V = 16 m/s and Hs = 5.5 m, although this environmental state Max. obs. 116.5 117.8 1.01 120.9 140.1 1.16
requires a rarer (3.87 × 10−6 ) probability of the exceedance of a Notes: Min, Max and Median are statistics of extrapolated loads over 10 sets. Max.
load given environment. obs. is the largest observed load from 500 simulations.

5.2. Convergence check for rare fractiles in Table 5, which are estimated using bootstrap resamplings, are
all smaller than the 82% confidence bounds shown in Table 4,
While we have performed 500 10-min simulations, resulting in which were estimated from non-repeating sets of real data (albeit
500 load extremes for these two environmental states, in practice, only ten sets). This is also expected as bootstrap resamplings
a much smaller number of simulations is typically carried out for (with replacement) give smaller confidence bounds than real non-
extrapolation. With a small number of simulations, however, the repeating random samples [46]. In summary, the stable estimates
accuracy of tails of short-term distribution and of rare fractile of a rare fractile and associated acceptable confidence bounds seen
estimates becomes questionable. In a previous study [27], we here suggest that 50 simulations are adequate for extrapolation.
proposed a convergence criterion, which compared confidence We now discuss extrapolated 20 year tower loads for the offshore
bounds (e.g., 90%) on an observed rare percentile (e.g., the 84th wind turbine being studied.
percentile) of the short-term load distribution versus an imposed
acceptable limit, to establish how many simulations would be 5.3. Extrapolated loads
adequate in order to have a stable distribution especially in tails.
Table 4 shows estimates of the 84th percentile 10-min maximum Table 6 shows extrapolated 20-year loads (resulting from
load from ten separate sets of 50 simulations each. The average of estimates of p3 fractiles). Listed are the minimum, maximum and
the ten estimates, shown in Table 4, is very close to the estimate of median estimates of long-term loads obtained from ten sets of 50
the 84th percentile load from 500 simulations, which is 100 MN m simulations. To estimate the p3 fractile, a two-parameter Weibull
(with linear waves) and 100.7 MN m (with nonlinear waves) for distribution was fitted to the above-median (i.e., the largest 25
V = 16 m/s, Hs = 5.5 m; and 96.1 MN m (with linear waves) and load extremes out of 50) load extremes. The required exceedance
108.3 MN m (with nonlinear waves) for V = 18 m/s, Hs = 7.5 m. probabilities (= 1 − p3 ) are 2.03 × 10−3 and 3.87 × 10−6 for V =
This suggests that 50 simulations may be sufficient to yield a stable 18 m/s, Hs = 7.5 m and V = 16 m/s, Hs = 5.5 m, respectively. It
estimate of a rare load fractile. It can also be seen that the 84th can be seen that the extrapolated loads are larger when nonlinear
percentile load estimate for all the cases is very stable, and it has waves are modeled, with the difference being about 18% for V =
a range (as defined in Table 4) that is always less than 10%. This 18 m/s, Hs = 7.5 m. The range of predicted loads (= 100 ×(Max −
range on the 84th percentile loads, which is essentially the 82% Min)/Median) is less than 15% for all cases. As was the case for the
(10/11 − 1/11 = 9/11 = 0.82) confidence interval relative to range on 84th percentile load in Table 4, the computed range in
the median estimate, is larger for nonlinear waves for both the Table 6 is larger for nonlinear waves, indicating greater variability
environmental states. This is expected since nonlinear waves are with nonlinear waves. Also note that the extrapolated loads are
more non-Gaussian and highly variable, and hence result in larger very close to the maximum observed load (over 500 simulations)
variability in turbine loads as well. for V = 18 m/s, Hs = 7.5 m; this is because the largest load
The 90% relative confidence bounds on the 84th percentile extreme from 500 simulations is associated with an exceedance
load, obtained using bootstrap techniques [46], are summarized in probability of 1/501 or 1.996 × 10−3 , which is close to the desired
Table 5. The relative confidence bounds (averaged over 10 sets) (1 − p3 ) exceedance probability of 2.03 × 10−3 .
from 50 simulations, are less than 10% for both environmental An important observation here is that the governing environ-
states, with linear and nonlinear waves. In an earlier paper [27], mental state changed from V = 16 m/s and Hs = 5.5 m when
we suggested that a given number of simulations is adequate to linear waves are modeled to V = 18 m/s and Hs = 7.5 m when
have a stable short-term distribution if such a relative confidence nonlinear waves are modeled. Note that we performed simulations
bound is less than 15%, which is the case here. Moreover, we see in for all combinations of wind speeds ranging from 3 to 25 m/s and
Table 5 that the relative confidence bound decreases significantly for wave heights ranging from 0 m to 10 m, both for linear and
when all 500 simulations are used in a similar bootstrapping nonlinear wave models, and we then used the inverse first-order
exercise. Besides, the 90% relative confidence bounds presented reliability method (FORM) to derive these governing states and the
P. Agarwal, L. Manuel / Applied Ocean Research 33 (2011) 215–227 225

a b

c d

Fig. 7. Scatter plots showing whether the occurrences of the maximum fore-aft tower bending moment in 10 minutes and the occurrence of the maximum instantaneous
wave height are close (i.e., well correlated) or far apart (i.e., less correlated). Plots are for V = 16 m/s and Hs = 5.5 m.

associated 20-year long-term loads. If we had ‘corrected’ long-term to those with linear waves. Furthermore, the scatter plot shows
loads by performing simulations with nonlinear waves only for the strong correlation between maximum wave heights and maximum
governing environmental state (V = 16 m/s, Hs = 5.5 m) for loads for the nonlinear wave model only for V = 18 m/s and
the linear waves, the long-term load would be 128.1 MN m. This is Hs = 7.5 m. This clearly indicates that nonlinear waves, especially
clearly smaller and, hence, non-conservative compared to the long- for large wave heights, have a clear influence on loads. Note that
term load of 138.4 MN m found after independently estimating the this is not the case for linear waves.
governing environmental state for nonlinear waves. In summary, the governing environmental state changed from
V = 16 m/s and Hs = 5.5 m when linear waves are modeled to
5.4. Correlation of occurrences of large wave heights and large loads V = 18 m/s and Hs = 7.5 m when nonlinear waves are modeled,
and the long-term load increased by about 10%. This is because at
To further understand how waves influence loads for the larger wave heights, wave loads, relative to wind loads, become
two environmental states studied, we investigate whether the more important when a nonlinear wave model is used. Therefore,
occurrence of the maximum fore-aft tower bending moment in estimating long-term loads for an offshore wind turbine sited
in a 10-min time series and the occurrence of the maximum in shallow water depths, it is important that nonlinear irregular
instantaneous wave height in that same time series are close waves are modeled.
(i.e., well correlated) or far apart (i.e., less correlated) in time. A
close occurrence would indicate that large waves might be causing 6. Summary and conclusions
large loads, for example in Fig. 4a. We choose the time separation
(to distinguish well correlated from less correlated) to be equal to Coupled simulation of wind turbines is a challenging problem
the peak spectral period of the waves. Scatter plots of well and which requires proper modeling and interfacing of very different
less correlated events in each set are shown in Figs. 7 and 8 for physics representing various parts of a wind turbine, namely,
V = 16 m/s and Hs = 5.5 m and for V = 18 m/s and Hs = 7.5 m, aerodynamics, aeroelasticity, controls, and hydrodynamics. We
respectively; the total number of such events are also indicated focussed on hydrodynamics and our objective was to improve
on each of the plots. It is seen that for V = 16 m/s and Hs = the accuracy of the coupled simulation by including a nonlinear
5.5 m, the fraction of well correlated events is about 9% and 14% irregular wave model that is more appropriate for shallow
for linear and nonlinear waves, respectively. For V = 18 m/s and water depths, where most offshore wind turbines are sited. We
Hs = 7.5 m, the fraction of well correlated events is higher — about incorporated a second-order nonlinear wave model in the coupled
21% and 44% for linear and nonlinear waves, respectively. In both aero-servo-hydro-elastic simulation of offshore wind turbines.
cases, the fraction of well correlated events is seen to get larger for We presented the theoretical formulation for the second-order
nonlinear waves. For V = 18 m/s and Hs = 7.5 m, this fraction nonlinear irregular waves as developed by Sharma and Dean [10].
is almost doubled, which is consistent with our earlier observation We investigated the effect of linear versus nonlinear irregular
that nonlinear waves for this environmental state result in larger wave models on the response and the long-term loads for a
wave heights and, consequently, larger loads. For V = 18 m/s utility-scale 5 MW offshore wind turbine model sited in 20 m of
and Hs = 7.5 m, it can also be seen from the scatter plots water. This turbine has a monopile support structure (a cylinder of
(Fig. 8) that larger loads occur with nonlinear waves, compared 6 m diameter). Our focus was primarily on the fore-aft tower
226 P. Agarwal, L. Manuel / Applied Ocean Research 33 (2011) 215–227

a b

c d

Fig. 8. Scatter plots showing whether the occurrences of the maximum fore-aft tower bending moment in 10 minutes and the occurrence of the maximum instantaneous
wave height are close (i.e., well correlated) or far apart (i.e., less correlated). Plots are for V = 18 m/s and Hs = 7.5 m.

bending moment at the mudline. We showed that for a given In summary, this study suggests that nonlinear irregular waves
environmental state, the nonlinear waves result in larger loads have an important influence on the long-term loads on an offshore
than those due to linear waves. The loads due to nonlinear waves wind turbine support structure, even when the turbine is in
tend to be more non-Gaussian and, therefore, extreme loads due operation when loads due to wind are generally much larger than
to nonlinear waves are predicted to be larger. Moreover, nonlinear those due to waves. Therefore, nonlinear irregular waves should
waves also influence the dynamic behavior of the tower loads. be considered when predicting long-term loads for offshore wind
We studied the influence of the alternate irregular wave models turbine design.
on the short-term load distributions, and on the long-term loads,
which are of interest in the design of turbines. We performed Acknowledgements
simulations of turbine response for all possible combinations
of wind speed and significant wave height, both for linear and The authors gratefully acknowledge assistance received with
nonlinear wave models. We used the inverse first-order reliability the 5MW offshore wind turbine baseline model used in the sim-
method (FORM) to compute 20-year long-term loads. We found ulation studies from Dr. Jason Jonkman of the National Renew-
that the long-term loads increased by about 10% when nonlinear able Energy Laboratory. The authors also wish to acknowledge
waves were modeled. Therefore, a design can be unconservative the financial support provided by the National Science Foundation
if only linear waves have been considered. More importantly, (CAREER Award No. CMMI-0449128 and Award No. CMMI-
the governing environmental state changed from V = 16 m/s 0727989) and by Sandia National Laboratories (Contract No.
and Hs = 5.5 m when linear waves are modeled to V = 743358).
18 m/s and Hs = 7.5 m when nonlinear waves are modeled. The
relative contribution of waves to wind towards overall long-term Appendix. Transfer functions for second-order irregular waves
loads increased at this latter environmental state. The approach
of finding the governing environmental state and associated The transfer functions B− +
mn and Bmn in Eq. (3) are given by:
long-term loads by performing coupled simulations for nonlinear
mn − (km kn + Rm Rn )
D−
waves is more accurate than an approach in which the loads for
[ ]
1
nonlinear waves are obtained only for governing environmental B−
mn = √ + (Rm + Rn ) (A.1)
4 Rm Rn
state corresponding to the linear wave model.
mn − (km kn − Rm Rn )
D+
[ ]
Furthermore, using correlations of occurrences of large loads 1
and large wave heights for the governing environmental state, we
B+
mn = √ + (Rm + Rn ) (A.2)
4 Rm Rn
showed that the influence of the nonlinear wave model becomes
clearly more pronounced at larger wave heights. This suggests where
√ √ √ √
( Rm − Rn ) Rn (k2m − R2m ) − Rm (k2n − R2n )

that modeling of nonlinear waves may be more important at
sites where waves are higher. Besides, we showed that confidence D−
mn = √ √ 2
( Rm − Rn ) − k− mn tanh(kmn h)

intervals on rare fractiles estimated from short-term distributions √ √ 2
are within acceptable limits, which establishes a reasonable 2( Rm − Rn ) (km kn + Rm Rn )
accuracy of long-term load predictions.
+ √ √ (A.3)
( Rm − Rn )2 − k−
mn tanh(kmn h)

P. Agarwal, L. Manuel / Applied Ocean Research 33 (2011) 215–227 227
√ √ √
( Rm + Rn ) ( − )+ ( − )
√ 
Rn k2m R2m Rm k2n R2n [19] Forristall GZ. Irregular wave kinematics from a kinematic boundary condition
D+
mn = √ √ 2 fit (KBCF). Applied Ocean Research 1985;7(4):202–12.
( Rm + Rn ) − kmn tanh(kmn h)
+ +
[20] Tromans PS, Anaturk AR, Hagemeijer P. A new model for kinematics of large
√ √ ocean waves — application as a design wave. In: 1st International offshore and
2( Rm + Rn )2 (km kn − Rm Rn )
+ √ √ . (A.4) polar engineering conference. vol. 3. 1991.
( Rm + Rn )2 − k+mn tanh(kmn h)
+ [21] Taylor PH, Jonathan P, Harland LA. Time-domain simulation of jack-up
dynamics with the extremes of a gaussian process. In: 14th International
In the above, k is the wave number which is related the conference on offshore mechanics and arctic engineering. vol. 1A. 1995.
p. 313–9.
frequency, ω, and the water depth, h, via the dispersion relation. [22] Zhang J, Chen L, Ye M, Randall RE. Hybrid wave model for unidirectional
Related parameters that are needed are given as follows: irregular waves — Part I. theory and numerical scheme. Appl Ocean Res 1996;
18:77–92.
ω2 = gk tanh(kh) (A.5) [23] Nwogu O. Alternative form of Boussinesq equations for nearshore wave
propagation. Journal of Waterway, Port, Coastal and Ocean Engineering 1993;
ω 2
m
119(6):618–38.
Rm = (A.6) [24] Madsen PA, Bingham HB, Liu H. A new Boussinesq method for fully nonlinear
g waves from shallow to deep water. Journal of Fluid Mechanics 2002;462:1–30.
[25] Jonkman JM, Buhl ML Jr. FAST user’s guide. Tech. rep. NREL/EL-500-38230;
k−
mn = |km − kn | (A.7) Golden (CO): National Renewable Energy Laboratory; 2005.
+ [26] Jonkman JM, Butterfield S, Musial W, Scott G. Definition of a 5-MW reference
kmn = km + kn (A.8) wind turbine for offshore system development. Tech. rep. NREL/TP-500-
38060; Golden (CO): National Renewable Energy Laboratory; 2007 [in press].
where g refers to acceleration due to gravity. [27] Fogle J, Agarwal P, Manuel L. Towards an improved understanding of statistical
extrapolation for wind turbine extreme loads. In: Wind energy. Design load
definition 2008;11(6):613–35 [special issue].
References [28] Burton T, Sharpe D, Jenkins N, Bossanyi E. Wind energy handbook. Chichester
(England): John Wiley; 2001.
[1] Musial W, Butterfield S. Future for offshore wind energy in the united states. [29] Barltrop NDP, Adams AJ. Dynamics of fixed marine structures. 3rd ed. London:
Tech. Rep. NREL/CP-500-36313; Golden (CO): National Renewable Energy Butterworth-Heinemann; 1991.
Laboratory; 2004. [30] Sarpkaya T, Issacson M. Mechanics of wave forces on offshore structures. New
[2] European Wind Energy Association. Response to the European commissions’s York: Van Nostrand Reinhold; 1981.
green paper: a European strategy for sustainable, competitive and secure [31] Sharma JN. Development and evaluation of a procedure for simulating a
energy, EWEA position paper. Brussels, Belgium; 2006. random directional second-order sea surface and associated wave forces. Ph.D.
[3] European Wind Energy Association. Delivering offshore wind power in Europe thesis. Delaware: University of Delaware; 1979.
— policy recommendations for large-scale deployment of offshore wind power [32] Hudspeth R, Chen MC. Digital simulation of nonlinear random waves. Journal
in Europe by 2020. Brussels, Belgium; 2007. of Waterway, Port, Coastal and Ocean Division, ASCE 1979;105(WW1):67–85.
[4] IEC-61400-3. Wind turbines—part 3: Design requirements for offshore wind [33] Hu SLJ. Nonlinear random water wave. In: Chang AH-D, Yang CY, editors.
turbines. Edition 1.0, International Electrotechnical Commission, 2009. Computational stochastic mechanics. Southampton (UK): Computational
[5] Saranyasoontorn K, Manuel L. Design loads for wind turbines using Mechanics Publication; 1993. p. 519–44.
the environmental contour method. Journal of Solar Energy Engineering, [34] Shinozuka M, Deodatis G. Simulation of stochastic processes by spectral
Transactions of the ASME 2006;128(4):554–61. representation. Applied Mechanics Reviews 1991;44(4):191–203.
[6] DNV-RP-C205. Environmental conditions and environmental loads, recom- [35] Massel SR. Ocean surface waves: their physics and prediction. Singapore:
mended practice. Det Norske Veritas; 2007. World Scientific Publishing; 1996.
[7] Forristall GZ. Nonlinear wave calculations for engineering applications. Journal [36] Stansberg CT, Gudmestad OT, Haver SK. Kinematics under extreme waves.
of Offshore Mechanics and Arctic Engineering 2002;124(1):28–33. Journal of Offshore Mechanics and Arctic Engineering 2008;130(2):
[8] Jha AK. Nonlinear stochastic models for ocean wave loads and responses 021010–1–021010–7.
of offshore structures and vessels. Ph.D. Dissertation. Stanford University; [37] Agarwal P, Manuel L. Wave models for offshore wind turbines. In: ASME wind
1997. energy symposium. Reno (NV): AIAA; 2008.
[9] ITTC. The specialist committee on environmental modeling, final report and [38] Jonkman BJ, Buhl ML Jr.. TurbSim User’s Guide. Tech. rep. NREL/TP-500-41136;
recommendations to the 22nd ittc. Tech. rep.; International towing tank Golden (CO): National Renewable Energy Laboratory; 2007.
conference; 1999. [39] DNV-OS-J101. Design of offshore wind turbine structures, offshore standard.
[10] Sharma JN, Dean RG. Development and evaluation of a procedure for Det Norske Veritas; 2007.
simulating a random directional second-order sea surface and associated wave [40] Sharma JN, Dean RG. Contributions to second order directional sea simulation
forces. Tech. rep. ocean engineering report no. 20; Newark (DE): University of and wave forces. In: 26th International conference on offshore mechanics and
Delaware; 1979. arctic engineering. Paper no. 29634; 2007.
[11] Longuet-Higgins M. Resonant interactions between two trains of gravity [41] Agarwal P, Manuel L. On the modeling of nonlinear waves for prediction of
waves. Journal of Fluid Mechanics 1962;12(3):321–32. long-term offshore wind turbine loads. In: 27th International conference on
[12] Hasselmann K. On the non-linear energy transfer in a gravity wave spectrum, offshore mechanics and arctic engineering. Paper no. 57855; 2008.
part 1, general theory. Journal of Fluid Mechanics 1962;12:481–500. [42] Jonkman J, Butterfield SP, Passon P, Larsen T, Camp T, Nichols J. et al. Offshore
[13] Hu SLJ, Zhao D. Non-gaussian properties of second-order random waves. code comparison collaboration within IEA wind Annex XXIII: Phase II results
Journal of Engineering Mechanics 1993;119(2):344–64. regarding monopile foundation modeling. Tech. rep. NREL/CP-500-42471;
[14] Langley R. A statistical analysis of nonlinear random waves. Ocean Engineering Golden (CO): National Renewable Energy Laboratory; 2008.
1987;14(5):389–407. [43] Tempel J, Molenaar DP. Wind turbine structural dynamics-a review of the
[15] Longuet-Higgins M. The propagation of short surface waves on longer gravity principles for modern power generation, onshore and offshore. Wind Eng
waves. Journal of Fluid Mechanics 1987;177:293–306. 2002;26(4):211–20.
[16] Forristall GZ. Wave crest distributions: observations and second-order theory. [44] Winterstein SR, Ude TC, Cornell CA, Bjerager P, Haver S. Environmental
Journal of Physical Oceanography 2000;30(8):1931–43. contours for extreme response: Inverse FORM with omission factors. In:
[17] Marthinsen T, Winterstein SR. On the skewness of random surface waves. In: Proceedings, ICOSSAR-93; 1993.
2nd International offshore and polar engineering conference; vol. 3. 1992. [45] Agarwal P, Manuel L. Simulation of offshore wind turbine response for long-
[18] Stansberg CT. Non-gaussian properties of second-order sum frequency term ultimate loads. Engineering Structures 2009;31(10):2236–46.
responses in irregular waves. In: 12th International conference on offshore [46] Efron B, Tibshirani RJ. An introduction to the bootstrap. New York: Chapman
mechanics and arctic engineering; vol. 1. 1993. and Hall; 1993.

You might also like