Download as pdf or txt
Download as pdf or txt
You are on page 1of 88

GH Bladed

Theory Manual
Document No
Classification
Issue no.
Date

282/BR/009
Commercial in Confidence
11
July 2003

Author:
E A Bossanyi
Checked by:
D C Quarton
Approved by:
D C Quarton

DISCLAIMER
Acceptance of this document by the client is on the basis that Garrad Hassan and
Partners Limited are not in any way to be held responsible for the application or use
made of the findings of the results from the analysis and that such responsibility
remains with the client.

Key To Document Classification


Strictly Confidential

Recipients only

Private and Confidential

For disclosure to individuals directly


concerned
within
the
recipients
organisation

Commercial in Confidence

Not to be disclosed outside the recipients


organisation

GHP only

Not to be disclosed to non GHP staff

Clients Discretion

Distribution at the discretion of the client


subject to contractual agreement

Published

Available to the general public

2003 Garrad Hassan and Partners Limited

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

CONTENTS

1.

Introduction
1.1
1.2
1.3
1.4
1.5

2.

Purpose
Theoretical background
Support
Documentation
Acknowledgements

1
2
3
3
3

AERODYNAMICS

2.1

Combined blade element and momentum theory


2.1.1 Actuator disk model
2.1.2 Wake rotation
2.1.3 Blade element theory
2.1.4 Tip and hub loss models
2.2
Wake models
2.2.1 Equilibrium wake
2.2.2 Frozen wake
2.2.3 Dynamic wake
2.3
Steady stall
2.4
Dynamic stall

3.

STRUCTURAL DYNAMICS

13

3.1

Modal analysis
3.1.1 Rotor modes
3.1.2 Tower modes
3.2
Equations of motion
3.2.1 Degrees of freedom
3.2.2 Formulation of equations of motion
3.2.3 Solution of the equations of motion
3.3
Calculation of structural loads

4.
4.1

4.2

4.3
4.4
4.5

5.
5.1
5.2

4
4
5
6
8
9
9
9
9
11
11
13
14
15
16
16
16
17
18

POWER TRAIN DYNAMICS

19

Drive train models


4.1.1 Locked speed model
4.1.2 Rigid shaft model
4.1.3 Flexible shaft model
Generator models
4.2.1 Fixed speed induction generator
4.2.2 Fixed speed induction generator: electrical model
4.2.3 Variable speed generator
4.2.4 Variable slip generator
Drive train mounting
Energy losses
The electrical network

19
19
19
19
20
20
21
22
23
24
24
25

CLOSED LOOP CONTROL

27

Introduction
The fixed speed pitch regulated controller

-i-

27
27

Garrad Hassan and Partners Ltd

Document: 282/BR/009

5.2.1 Steady state parameters


5.2.2 Dynamic parameters
5.3
The variable speed stall regulated controller
5.3.1 Steady state parameters
5.3.2 Dynamic parameters
5.4
The variable speed pitch regulated controller
5.4.1 Steady state parameters
5.4.2 Dynamic parameters
5.5
Transducer models
5.6
Modelling the pitch actuator
5.7
The PI control algorithm
5.7.1 Gain scheduling
5.8
Control mode changes
5.9
Client-specific controllers
5.10 Signal noise and discretisation

6.

SUPERVISORY CONTROL

ISSUE:011

FINAL

28
28
28
28
30
31
31
32
33
33
36
37
38
38
39

40

6.1
6.2
6.3
6.4
6.5
6.6

Start-up
Normal stops
Emergency stops
Brake dynamics
Idling and parked simulations
Yaw control
6.6.1 Active yaw
6.6.2 Yaw dynamics
6.7
Teeter restraint

40
41
41
42
42
42
42
43
44

MODELLING THE WIND

45

Wind shear
7.1.1 Exponential model
7.1.2 Logarithmic model
Tower shadow
7.2.1 Potential flow model
7.2.2 Empirical model
7.2.3 Combined model
Upwind turbine wake
7.3.1 Eddy viscosity model of the upwind turbine wake
7.3.2 Turbulence in the wake
Time varying wind
7.4.1 Single point time history
7.4.2 3D turbulent wind
7.4.3 IEC transients
Three dimensional turbulence model
7.5.1 The basic von Karman model
7.5.2 The improved von Karman model
7.5.3 The Kaimal model
7.5.4 Compatibility with IEC 1400-1
7.5.5 Using 3d turbulent wind fields in simulations

46
46
46
46
46
47
47
47
48
50
51
51
51
52
53
53
55
59
59
59

MODELLING WAVES AND CURRENTS

61

7.
7.1
7.2

7.3
7.4

7.5

8.
8.1
8.2

Tower and Foundation Model


Wave Spectra

61
62
- ii -

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

8.2.1 JONSWAP / Pierson-Moskowitz Spectrum


8.2.2 User-defined Spectrum
8.3
Upper Frequency Limit
8.4
Wave Particle Kinematics
8.5
Wheeler Stretching
8.6
Simulation of Irregular Waves
8.7
Simulation of Regular Waves
8.8
Current Velocities
8.8.1 Near-Surface Current
8.8.2 Sub-Surface Current
8.8.3 Near-Shore Current
8.9
Total Velocities and Accelerations
8.10 Applied Forces
8.10.1
Relative Motion Form of Morisons Equation
8.10.2
Longitudinal Pressure Forces on Cylindrical Elements

9.

POST-PROCESSING

62
62
63
63
64
64
66
67
68
68
68
69
69
69
69

71

9.1
9.2
9.3
9.4
9.5
9.6

Basic statistics
Fourier harmonics, and periodic and stochastic components
Extreme prediction
Spectral analysis
Probability, peak and level crossing analysis
Rainflow cycle counting and fatigue analysis
9.6.1 Rainflow cycle counting
9.6.2 Fatigue analysis
9.7
Annual energy yield
9.8
Ultimate loads
9.9
Flicker

10.

FINAL

References

71
71
72
75
75
76
76
77
78
79
79

80

- iii -

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

1. INTRODUCTION

1.1 Purpose
GH Bladed is an integrated software package for wind turbine performance and loading
calculations. It is intended for the following applications:
Preliminary wind turbine design
Detailed design and component specification
Certification of wind turbines
With its sophisticated graphical user interface, it allows the user to carry out the following
tasks in a straightforward way:
Specification of all wind turbine parameters, wind inputs and load cases.
Rapid calculation of steady-state performance characteristics, including:
Aerodynamic information
Performance coefficients
Power curves
Steady operating loads
Steady parked loads
Dynamic simulations covering the following cases:
Normal running
Start-up
Normal and emergency shut-downs
Idling
Parked
Dynamic power curve
Post-processing of results to obtain:
Basic statistics
Periodic component analysis
Probability density, peak value and level crossing analysis
Spectral analysis
Cross-spectrum, coherence and transfer function analysis
Rainflow cycle counting and fatigue analysis
Combinations of variables
Annual energy yield
Ultimate loads (identification of worst cases)
Flicker severity
Presentation: results may be presented graphically and can be combined into a word
processor compatible report.

1 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

1.2 Theoretical background


The Garrad Hassan approach to the calculation of wind turbine performance and loading has
been developed over many years. The main aim of this development has been to produce
reliable tools for use in the design and certification of wind turbines.
The models and theoretical methods incorporated in GH Bladed have been extensively
validated against monitored data from a wide range of turbines of many different sizes and
configurations, including:

WEG MS-1, UK, 1991


Howden HWP300 and HWP330, USA, 1993
ECN 25m HAT, Netherlands, 1993
Newinco 500kW, Netherlands, 1993
Nordex 26m, Denmark, 1993
Nibe A, Denmark, 1993
Holec WPS30, Netherlands, 1993
Riva Calzoni M30, Italy, 1993
Nordtank 300kW, Denmark, 1994
WindMaster 750kW, Netherlands, 1994
Tjaereborg 2MW, Denmark, 1994
Zond Z-40, USA, 1994
Nordtank 500kW, UK, 1995
Vestas V27, Greece, 1995
Danwin 200kW, Sweden, 1995
Carter 300kW, UK, 1995
NedWind 50, 1MW, Netherlands, 1996
DESA, 300kW, Spain 1997
NTK 600, UK, 1998
West Medit, Italy, 1998
Nordex 1.3 MW, Germany, 1999
The Wind Turbine Company 350 kW, USA, 2000
Windtec 1.3 MW, Austria, 2000
WEG MS-4, 400 kW, UK, 2000
EHN 1.3 MW, Spain, 2001
Vestas 2MW, UK, 2001
Lagerwey 750 Netherlands, 2001
Vergnet 200, France 2001

This document describes the theoretical background to the various models and numerical
methods incorporated in GH Bladed.

2 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

1.3 Support
GH Bladed is supplied with a one-year maintenance and support agreement, which can be
renewed for further periods. This support includes a hot-line help service by telephone, fax
or e-mail:
Telephone:
Fax:
E-mail

+44 (0)117 972 9900


+44 (0)117 972 9901
bladed@bristol.garradhassan.co.uk

1.4 Documentation
In addition to this Theory Manual, there is also a GH Bladed User Manual which explains
how the code can be used.

1.5 Acknowledgements
GH Bladed was developed with assistance from the Commission of the European
Communities under the JOULE II programme, project no. JOU2-CT92-0198.

3 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

2. AERODYNAMICS
The modelling of rotor aerodynamics provided by Bladed is based on the well established
treatment of combined blade element and momentum theory [2.1]. Two major extensions of this
theory are provided as options in the code to deal with the unsteady nature of the aerodynamics.
The first of these extensions allows a treatment of the dynamics of the wake and the second
provides a representation of dynamic stall through the use of a stall hysteresis model.
The theoretical background to the various aspects of the treatment of rotor aerodynamics
provided by Bladed is given in the following sections.

2.1 Combined blade element and momentum theory


At the core of the aerodynamic model provided by Bladed is combined blade element and
momentum theory. The features of this treatment of rotor aerodynamics are described below.
2.1.1 Actuator disk model
To aid the understanding of combined blade element and momentum theory it is useful initially
to consider the rotor as an actuator disk. Although this model is very simple, it does provide
valuable insight into the aerodynamics of the rotor.
Wind turbines extract energy from the wind by producing a step change in static pressure across
the rotor-swept surface. As the air approaches the rotor it slows down gradually, resulting in an
increase in static pressure. The reduction in static pressure across the rotor disk results in the air
behind it being at sub atmospheric pressure. As the air proceeds downstream the pressure climbs
back to the atmospheric value resulting in a further slowing down of the wind. There is therefore
a reduction in the kinetic energy in the wind, some of which is converted into useful energy by
the turbine.
In the actuator disk model of the process described above, the wind velocity at the rotor disk Ud
is related to the upstream wind velocity Uo as follows:
U d = ( 1 a )U o

The reduced wind velocity at the rotor disk is clearly determined by the magnitude of a, the axial
flow induction factor or inflow factor.
By applying Bernoullis equation and assuming the flow to be uniform and incompressible, it
can be shown that the power P extracted by the rotor is given by :
P = 2 AU o3a( 1 a )3

where

is the air density and A the area of the rotor disk.

The thrust T acting on the rotor disk can similarly be derived to give:

4 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

T = 2 AU o2 a( 1 a )

The dimensionless power and thrust coefficients, CP and CT are respectively:


CP = P / ( 1 2 AU o3 ) = 4a( 1 a )2

and:
CT = T / ( 1 2 AU o2 ) = 4a( 1 a )

The maximum value of the power coefficient CP occurs when a is 1 /3 and is equal to 16/27 which
is known as the Betz limit.
The thrust coefficient CT has a maximum value of 1 when a is 1 /2.
2.1.2 Wake rotation
The actuator disk concept used above allows an estimate of the energy extracted from the wind
without considering that the power absorbed by the rotor is the product of torque Q and angular
velocity
of the rotor. The torque developed by the rotor must impart an equal and opposite
rate of change of angular momentum to the wind and therefore induces a tangential velocity to
the flow. The change in tangential velocity is expressed in terms of a tangential flow induction
factor a. Upstream of the rotor disk the tangential velocity is zero, at the disk the tangential
velocity at radius r on the rotor is ra and far downstream the tangential velocity is 2 ra.
Because it is produced in reaction to the torque, the tangential velocity is opposed to the motion
of the blades.
The torque generated by the rotor is equal to the rate of change of angular momentum and can be
derived as:
Q=

R 4 (1 a )a ,U o

5 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

2.1.3 Blade element theory


Combined blade element and momentum theory is an extension of the actuator disk theory
described above. The rotor blades are divided into a number of blade elements and the theory
outlined above used not for the rotor disk as a whole but for a series of annuli swept out by each
blade element and where each annulus is assumed to act in the same way as an independent
actuator disk. At each radial position the rate of change of axial and angular momentum are
equated with the thrust and torque produced by each blade element.
The thrust dT developed by a blade element of length dr located at a radius r is given by:
dT = 1 2 W 2 ( CL cos + CD sin )cdr

where W is the magnitude of the apparent wind speed vector at the blade element,
is
known as the inflow angle and defines the direction of the apparent wind speed vector
relative to the plane of rotation of the blade, c is the chord of the blade element and CL and
CD are the lift and drag coefficients respectively.
The lift and drag coefficients are defined for an aerofoil by:
CL = L / ( 1 2 V 2 S )

and
CD = D / ( 1 2 V 2 S )

where L and D are the lift and drag forces, S is the planform area of the aerofoil and V is the
wind velocity relative to the aerofoil.
The torque dQ developed by a blade element of length dr located at a radius r is given by:
dQ = 1 2 W 2 r( CL sin

CD cos )cdr

In order to solve for the axial and tangential flow induction factors appropriate to the radial
position of a particular blade element, the thrust and torque developed by the element are
equated to the rate of change of axial and angular momentum through the annulus swept out
by the element. Using expressions for the axial and angular momentum similar to those
derived for the actuator disk in Sections 2.1.1 and 2.1.2 above, the annular induction factors
may be expressed as follows:
a = g1 / ( 1 + g1 )

and
a , = g2 / ( 1 g 2 )

where
6 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

g1 =

Bc ( CL cos + CD sin )
H
2 r
4 F sin 2

g2 =

Bc ( CL sin
CD cos )
2 r
4 F sin cos

ISSUE:011

FINAL

and

Here B is the number of blades and F is a factor to take account of tip and hub losses, refer
Section 2.1.4.
The parameter H is defined as follows:
for a

0.3539, H = 10
.

for a > 0.3539, H =

4a (1 a )
(0.6 + 0.61a + 0.79a 2 )

In the situation where the axial induction factor a is greater than 0.5, the rotor is heavily
loaded and operating in what is referred to as the turbulent wake state. Under these
conditions the actuator disk theory presented in Section 2.1.1 is no longer valid and the
expression derived for the thrust coefficient:
CT = 4a( 1 a )

must be replaced by the empirical expression:


CT = 0.6 + 0.61a + 0.79a 2

The implementation of blade element theory in Bladed is based on a transition to the


empirical model for values of a greater than 0.3539 rather than 0.5. This strategy results in a
smoother transition between the models of the two flow states.
The equations presented above for a and a can only be solved iteratively. The procedure
involves making an initial estimate of a and a, calculating the parameters g1 and g2 as
functions of a and a, and then using the equations above to update the values of a and a.
This procedure continues until a and a have converged on a solution. In Bladed convergence
is assumed to have occurred when:
ak

ak

tol

a'k

a'k

tol

and

where tol is the value of aerodynamic tolerance specified by the user.

7 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

2.1.4 Tip and hub loss models


The wake of the wind turbine rotor is made up of helical sheets of vorticity trailed from each
rotor blade. As a result the induced velocities at a fixed point on the rotor disk are not constant
with time, but fluctuate between the passage of each blade. The greater the pitch of the helical
sheets and the fewer the number of blades, the greater the amplitude of the variation of induced
velocities. The overall effect is to reduce the net momentum change and so reduce the net power
extracted. If the induction factor a is defined as being the value which applies at the instant a
blade passes a given point on the disk, then the average induction factor at that point, over the
course of one revolution will be aFt,, where Ft is a factor which is less than unity.
The circulation at the blade tips is reduced to zero by the wake vorticity in the same manner as at
the tips of an aircraft wing. At the tips, therefore the factor Ft becomes zero. Because of the
analogy with the aircraft wing , where losses are caused by the vortices trailing from the tips, Ft
is known as the tip loss factor.
Prandtl [2.2] put forward a method to deal with this effect in propeller theory. Reasoning that, in
the far wake, the helical vortex sheets could be replaced by solid disks, set at the same pitch as
the normal spacing between successive turns of the sheets, moving downstream with the speed
of the wake.
The flow velocity outside of the wake is the free stream value and so is faster than that of the
disks. At the edges of the disks the fast moving free stream flow weaves in and out between
them and in doing so causes the mean axial velocity between the disks to be higher than that of
the disks themselves, thus simulating the reduction in the change of momentum.
The factor Ft can be expressed in closed solution form:
Ft = 2 arccos[exp(

s )]
d

where s is the distance of the radial station from the tip of the rotor blade and d is the distance
between successive helical sheets.
A similar loss takes place at the blade root where, as at the tip, the bound circulation must fall to
zero and therefore a vortex must be trailed into the wake, A separate hub loss factor Fh is
therefore calculated and the effective total loss factor at any station on the blade is then the
product of the two:
F = Ft Fh

The combined tip and hub loss factor is incorporated in the equations of blade element
theory as indicated in Section 2.1.3 above.

8 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

2.2 Wake models


2.2.1 Equilibrium wake
The use of blade element theory for time domain dynamic simulations of wind turbine
behaviour has traditionally been based on the assumption that the wake reacts instantaneously
to changes in blade loading. This treatment, known as an equilibrium wake model, involves a
re-calculation of the axial and tangential induction factors at each element of each rotor
blade, and at each time step of a dynamic simulation. Based on this treatment the induced
velocities along each blade are computed as instantaneous solutions to the particular flow
conditions and loading experienced by each element of each blade.
Clearly in this interpretation of blade element theory the axial and tangential induced
velocities at a particular blade element vary with time and are not constant within the annulus
swept out by the element.
The equilibrium wake treatment of blade element theory is the most computationally
demanding of the three treatments described here.
2.2.2 Frozen wake
In the frozen wake model, the axial and tangential induced velocities are computed using
blade element theory for a uniform wind field at the mean hub height wind speed of the
simulated wind conditions. The induced velocities, computed according to the mean, uniform
flow conditions, are then assumed to be fixed, or frozen in time. The induced velocities
vary from one element to the next along the blade but are constant within the annulus swept
out by the element. As a consequence each blade experiences the same radial distribution of
induced flow..
It is important to note that it is the axial and tangential induced velocities aUo and ar
not the induction factors a and a which are frozen in time.

and

2.2.3 Dynamic wake


As described above, the equilibrium wake model assumes that the wake and therefore the
induced velocity flow field react instantaneously to changes in blade loading. On the other
hand, the frozen wake model assumes that induced flow field is completely independent of
changes in incident wind conditions and blade loading. In reality neither of these treatments
is strictly correct. Changes in blade loading change the vorticity that is trailed into the rotor
wake and the full effect of these changes takes a finite time to change the induced flow field.
The dynamics associated with this process is commonly referred to as dynamic inflow.
The study of dynamic inflow was initiated nearly 40 years ago in the context of helicopter
aerodynamics. In brief, the theory provides a means of describing the dynamic dependence of
the induced flow field at the rotor upon the loading that it experiences. The dynamic inflow
model used within Bladed is based on the work of Pitt and Peters [2.3] which has received
substantial validation in the helicopter field, see for example Gaonkar et al [2.4].
The Pitt and Peters model was originally developed for an actuator disk with assumptions
made concerning the distribution of inflow across the disc. In Bladed the model is applied at
blade element or actuator annuli level since this avoids any assumptions about the
distribution of inflow across the disc.
9 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

For a blade element, bounded by radii R1 and R2 , and subject to uniform axial flow at a wind
speed Uo, the elemental thrust, dT, can be expressed as:
dT = 2U o am + U o m A a&

where m is the mass flow through the annulus, mA is the apparent mass acted upon by the
annulus and a is the axial induction factor.
The mass flow through the annular element is given by:
m = U o (1 a )dA

where dA is the cross-sectional area of the annulus.


For a disc of radius R the apparent mass upon which it acts is given approximately by
potential theory, Tuckerman, [2.5]:
mA = 8

R3

Therefore the thrust coefficient associated with the annulus can be derived to give:
C T = 4a (1 a ) +

16 (R 32
3 U o (R 22

R 13 )
R 12 )

a&

This differential equation can therefore be used to replace the blade element and momentum
theory equation for the calculation of axial inflow. The equation is integrated at each time
step to give time dependent values of inflow for each blade element on each blade. The
tangential inflow is obtained in the usual manner and so depends on the time dependent axial
value. It is evident that the equation introduces a time lag into the calculation of inflow which
is dependent on the radial station.
It is probable that the values of time lag for each blade element calculated in this manner will
under-estimate somewhat the effects of dynamic inflow, as each element is treated
independently with no consideration of the three dimensional nature of the wake or the
possibly dominant effect of the tip vortex. The treatment is, however, consistent with blade
element theory and provides a simple, computationally inexpensive and reasonably reliable
method of modelling the dynamics of the rotor wake and induced velocity flow field.

10 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

2.3 Steady stall


The representation and to some extent the general understanding of aerodynamic stall on a
rotating wind turbine blade remain rather poor. This is a rather extraordinary situation in
view of the importance of stall regulation to the industry.
Stall delay on the inboard sections of rotor blades, due to the three dimensionality of the
incident flow field, has been widely confirmed by measurements at both model and full scale.
A number of semi-empirical models [2.6, 2.7] have been developed for correcting two
dimensional aerofoil data to account for stall delay. Although such models are used for the
design analysis of stall regulated rotors, their general validity for use with a wide range of
aerofoil sections and rotor configurations remains, at present, rather poor. As a consequence
Bladed does not incorporate models for the modification of aerofoil data to deal with stall
delay, but the user is clearly able to apply whatever correction of the aerofoil data he believes
is appropriate prior to its input to the code.

2.4 Dynamic stall


Stall and its consequences are fundamentally important to the design and operation of most
aerodynamic devices. Most conventional aeronautical applications avoid stall by operating
well below the static stall angle of any aerofoils used. Helicopters and stall regulated
wind turbines do however operate in regimes where at least part of their rotor blades are in
stall. Indeed stall regulated wind turbines rely on the stalling behaviour of aerofoils to limit
maximum power output from the rotor in high winds.
A certain degree of unsteadiness always accompanies the turbulent flow over an aerofoil
at high angles of attack. The stall of a lifting surface undergoing unsteady motion is more
complex than static stall.
On an oscillating aerofoil, where the incidence is increasing rapidly, the onset of the stall can
be delayed to an incidence considerably in excess of the static stall angle. When dynamic stall
does occur, however, it is usually more severe than static stall. The attendant aerodynamic
forces and moments exhibit large hysteresis with respect to the instantaneous angle of
attack, especially if the oscillation is about a mean angle close to the static stall angle. This
represents an important contrast to the quasi-steady case, for which the flow field adjusts
immediately, and uniquely, to each change in incidence.
Many methods of predicting the dynamic stall of aerofoil sections have been developed,
principally for use in the helicopter industry.
The model adopted for inclusion of unsteady behaviour of aerofoils is that due to
Beddoes [2.8]. The Beddoes model was developed for use in helicopter rotor performance
calculations and has been formulated over a number of years with particular reference to
dynamic wind tunnel testing of aerofoil sections used on helicopter rotors. It has been used
successfully by Harris [2.9] and Galbraith et al [2.10] in the prediction of the behaviour of
vertical axis wind turbines.

11 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

The model used within Bladed is a development of the Beddoes model which has been validated
against measurements from several stall regulated wind turbines. The model utilises the
following elements of the method described in [2.8] to calculate the unsteady lift coefficient
The indicial response functions for modelling of attached flow
The time lagged Kirchoff formulation for the modelling of trailing edge separation and
vortex lift
The use of the model of leading edge separation has been found to be inappropriate for use on
horizontal axis wind turbines where the aerofoil characteristics are dominated by progressive
trailing edge stall.
The time lag in the development of trailing edge separation is a user defined parameter within
the model implemented in Bladed. This time lag encompasses the delay in the response of the
pressure distribution and boundary layer to the time varying angle of attack. The magnitude of
the time lag is directly related to the level of hysteresis in the lift coefficient.
The drag and pitching moment coefficients are calculated using the quasi-steady input data along
with the effective unsteady angle of attack determined during the calculation of the lift
coefficient.

12 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

3. STRUCTURAL DYNAMICS
In the early days of the industry, wind turbine design was undertaken on the basis of quasistatic aerodynamic calculations with the effects of structural dynamics either ignored
completely or included through the use of estimated dynamic magnification factors. From the
late 1970s research workers began to consider more reliable methods of dynamic analysis
and two basic approaches were considered: finite element representations and modal analysis.
The traditional use of standard, commercial finite element analysis codes for dealing with
problems of structural dynamics is problematic in the case of wind turbines. This is because
of the gross movement of one component of the structure, the rotor, with respect to another,
the tower. Standard finite element packages are only used to consider structures in which
motion occurs about a mean undisplaced position and for this reason the finite element
models of wind turbines which have been developed have been specially constructed to deal
with the problem.
The form of wind turbine dynamic modelling most commonly used as the basis of design
calculations is that involving a modal representation. This approach, borrowed from the
helicopter industry, has the major advantage that it offers a reliable representation of the
dynamics of a wind turbine with relatively few degrees of freedom. The number and type of
modal degrees of freedom used to represent the dynamics of a particular wind turbine will
clearly depend on the configuration and structural properties of the machine.
At present, largely because of the very extensive computer processing requirements
associated with the use of finite element models, the state of the art in the context of wind
turbine dynamic modelling for design analysis is based squarely on the use of limited degree
of freedom modal models. The representation of wind turbine structural dynamics within
Bladed is based on a modal model.

3.1 Modal analysis


Because of the rotation of the blades of a wind turbine relative to the tower support structure,
the equations of motion which describe its dynamics contain terms with periodic coefficients.
This periodicity means that the computation of the modal properties of an operating wind
turbine as a complete structural entity is not possible using the standard eigen-analysis
offered by commercial finite element codes.
One solution to this problem is to make use of Floquet analysis to determine the modal
properties of the periodic system. However, the mode shapes obtained by such calculations
are complex and not directly useful for a forced response analysis.
An alternative solution is based on the use of component mode synthesis. Here the modal
properties of the rotating and non-rotating components of the wind turbine are computed
independently. The component modes are then coupled by an appropriate formulation of the
equations of motion of the wind turbine in the forced response analysis. This approach has
been adopted for Bladed.

13 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

3.1.1 Rotor modes


The vibration of the tapered and twisted blades of a wind turbine rotor is a complex
phenomenon. A classical method of representing the vibration is by means of the orthogonal,
uncoupled normal modes of the structure. Each mode is defined in terms of the following
parameters:
Modal frequency,

Modal damping coefficient,


Mode shape,

(r )

where the subscript i indicates properties related to the ith mode.


The modal frequencies and mode shapes of the rotor are calculated based on the following
information:
The mass distribution along the blade.
The mass distribution is defined as the local mass density (kg/m) at each radial station in
addition to the magnitude and location of any discrete, lumped masses.
The bending stiffnesses along the blade.
The bending stiffnesses are defined in local flapwise and edgewise directions at each radial
station.
The twist angle distribution along the blade.
The mode shapes are computed in the rotor in-plane and out-of-plane directions and hence
the flapwise and edgewise stiffnesses at each radial station are resolved through the local
twist angle.
The blade pitch and setting angles.
The mode shapes are computed in the rotor in-plane and out-of-plane directions and hence
the flapwise and edgewise stiffnesses at each radial station are resolved through the blade
pitch and setting angles. The user of Bladed may select a series of different pitch angles for
which the modal analysis is carried out. During subsequent dynamic simulations, the modal
frequencies appropriate to the instantaneous blade pitch angle are therefore obtained by linear
interpolation of the results of the modal analyses.
The presence or otherwise of a hub teeter hinge for a two bladed rotor.
For a two-bladed rotor the hub can be rigid or teetered. The presence of a teeter hinge will
introduce asymmetric rotor modes involving out-of-plane rotation of the rotor about the teeter
hinge.
The presence or otherwise of a flap hinge for a one-bladed rotor.
For a one-bladed rotor the hub can be rigid or have a flap hinge. The presence of a flap hinge
will introduce rotor modes involving out-of-plane rotation of the rotor about the teeter hinge.
The counter-weight mass and moment of inertia about the flap hinge for a one-bladed rotor.
Whether the hub can rotate.
14 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

Rotation of the hub will affect the frequencies and mode shapes of the in-plane rotor modes.
With the shaft brake engaged and the rotor locked in position, the in-plane modes will
include both symmetric and asymmetric cantilever-type modes. With the rotor free to rotate,
the cantilever-type asymmetric modes will be replaced by asymmetric modes involving
rotation about the rotor shaft.
The rotational speed of the rotor.
The frequencies and mode shapes of both in-plane and out-of-plane modes will be dependent
on the rotational speed of the rotor. This dependence is explained by the additional bending
stiffness developed because of centrifugal loads acting on the deflected rotor blades. The user
of Bladed may select different rotational speeds for which the modal analysis is carried out.
During subsequent dynamic simulations, the modal frequencies appropriate to the
instantaneous rotational speed are therefore obtained by quadratic interpolation of the results
of the modal analyses.
The frequencies and mode shapes of the rotor modes are computed from the eigen-values and
eigen-vectors of a finite element representation of the rotor structure. The finite element
model of the rotor is based on the use of two-dimensional beam elements to describe the mass
and stiffness properties of the rotor blades.
The outputs from the modal analysis of the rotor are the modal frequencies and mode shapes
defined in the rotor in-plane and out-of-plane directions. The modal damping coefficients are
an input defined by the user and may be used to represent structural damping.
3.1.2 Tower modes
The representation of the bending dynamics of the tower is based on the modal degrees of
freedom in the fore-aft and side-side directions of motion. As for the rotor, the tower modes
are defined in terms of their modal frequency, modal damping and mode shape.
The modal frequencies and mode shapes of the tower are calculated based on the following
information:
The mass distribution along the tower.
The mass distribution is defined as the local mass density (kg/m) at each tower station height
in addition to the magnitude and location of any discrete, lumped masses.
The bending stiffness along the tower.
The tower is assumed to be axisymmetric with the bending stiffness therefore independent of
bending direction.
The mass, inertia and stiffness properties of the tower foundation.
The influence of the foundation mass and stiffness properties on the tower bending modes
may be taken into account. The model takes account of motion of the foundation mass and
inertia against both translational and rotational stiffnesses.
The mass and inertia of the nacelle and rotor
For calculation of the tower modes, the nacelle and rotor are modelled as lumped mass and
inertia located at the nacelle centre of gravity and rotor hub respectively. For one and twobladed rotors, the influence of the rotor inertia on the tower modal characteristics depends on
the rotor azimuth and this may therefore be defined by the user. The variation of the tower
modal frequencies with rotor azimuth is normally small and the assumption of a single rotor

15 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

azimuthal position for the modal analysis is therefore a reasonable approximation. The user
can, of course, determine the extent of the azimuthal variation in the tower modal frequencies
by undertaking the modal analysis at a series of different rotor azimuths.
The frequencies and mode shapes of the tower modes are computed from the eigen-values
and eigen-vectors of a finite element representation of the tower structure. The finite element
model of the tower is based on the use of two-dimensional beam elements to describe the
mass and stiffness properties of the tower.
The outputs from the modal analysis of the tower are the modal frequencies and mode shapes
defined in the fore-aft and side-side directions. The modal damping coefficients are an input
defined by the user and may be used to represent structural damping.

3.2 Equations of motion


Because of the complexity of the coupling of the modal degrees of freedom of the rotating
and non-rotating components, the algebraic manipulation involved in the derivation of the
equations of motion for a wind turbine is a complicated problem. In the case of the dynamic
model within Bladed, the derivation has been carried out using energy principles and
Lagrange equations by means of a computer algebra package.
3.2.1 Degrees of freedom
The degrees of freedom involved in the equations of motion for the structural dynamic model
for Bladed are as follows:

Rotor out of plane including teeter, maximum six modes


Rotor in-plane, maximum six modes
Nacelle yaw
Tower fore-aft, maximum three modes
Tower side-side, maximum three modes

In addition, a sophisticated representation of the power train dynamics is offered as described


in Section 4 of this manual.
3.2.2 Formulation of equations of motion
The equation of motion for a single modal degree of freedom, assuming no coupling with
other degrees of freedom, is as follows:

q&&i + 2

q& +

i i

2
i

q = Fi / Mi

where:
qi is the time dependent modal displacement,

Mi =

m(r )

2
i

(r )dr is the modal mass,

rotor

16 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

and:

Fi =

f (r ) i (r )dr is the modal force.


rotor

Here f(r) is the distributed force over the rotor or tower component.
The modal degrees of freedom are, of course, coupled and the formulation of the equations of
motion within Bladed is as follows:
&& + [ C]q& + [ K ]q = F
[ M ]q

where [M], [C] and [K] are the modal mass, damping and stiffness matrices, q is the vector
of modal displacements and F the vector of modal forces. The system matrices are full due to
the coupling of the degrees of freedom and contain periodic coefficients because of the time
dependent interaction of the dynamics of the rotor and tower.
Because of their complexity, the equations of motion are not presented in this manual. The
following key comments are, however, provided:
Although the equations of motion are based on a linear modal treatment of the structural
dynamics, the model does contain non-linear terms associated primarily with gyroscopic
coupling.
The rotor teeter degree of freedom is provided through the first out-of-plane mode and the
equation of motion includes representation of mechanical damping, stiffness and pre-load
restraints as specified by the user.
The equation of motion for the nacelle yaw degree of freedom is based on the inertia of
the wind turbine about the yaw axis with mechanical restraints provided through yaw
damping and stiffness as specified by the user.
The aeroelasticity of the wind turbine is taken into account in the equations of motion by
consideration of the interaction of the total structural velocity vector with the wind
velocity vector at each element along the rotor blades. The total structural velocity vector
at each element on the rotor blades is composed of the appropriate summation of the
velocities associated with each structural degree of freedom. In addition to the feedback of
the structural velocities into the rotor blade aerodynamics, the structural displacement
associated with the rotor teeter and nacelle yaw is also taken into account.
3.2.3 Solution of the equations of motion
The equations of motion are solved by time-marching integration of the differential equations
using a variable step size, fourth order Runge Kutta integrator.

17 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

3.3 Calculation of structural loads


The structural loads acting on the rotor, power train and tower are computed by the
appropriate summation of the applied aerodynamic loads and the inertial loads. The inertial
loads are calculated by integration of the mass properties and the total acceleration vector at
each station. The total acceleration vector includes modal, centrifugal, Coriolis and
gravitational components.

18 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

4. POWER TRAIN DYNAMICS


The power train dynamics define the rotational degrees of freedom associated with the drive
train, including drive train mountings, and the dynamics of the electrical generator. The drive
train consists of a low speed shaft, gearbox and high speed shaft. Direct drive generators can
also be modelled.

4.1 Drive train models


4.1.1 Locked speed model
The simplest drive train model which is available is the locked speed model, which allows no
degrees of freedom for the power train. The rotor is therefore assumed to rotate at an
absolutely constant speed, and the aerodynamic torque is assumed to be exactly balanced by
the generator reaction torque at every instant. Clearly this model is unsuitable for start-up
and shut-down simulations, but it is useful for quick, preliminary calculations of loads and
performance before the drive train and generator have been fully characterised.
4.1.2 Rigid shaft model
The rigid shaft model is obtained by selecting the dynamic drive train model with no shaft
torsional flexibility. It allows a single rotational degree of freedom for the rotor and
generator. It can be used for all calculations and is recommended if the torsional stiffness of
the drive train is high. The acceleration of the generator and rotor are calculated from the
torque imbalance divided by the combined inertia of the rotor and generator, making
allowance for the gearbox ratio. Direct drive generators are modelled simply by setting the
gearbox ratio to 1. The torque imbalance is essentially the difference between the
aerodynamic torque and the generator reaction torque and any applied brake torque, taking
the gearbox ratio into account. However, this is corrected to account for the inertial effect of
blade deflection due to any edgewise blade vibration modes. To use the rigid shaft model, a
model of the generator must also be provided, so that the generator reaction torque is defined.
During a parked simulation, or once the brake has brought the rotor to rest during a stopping
simulation, the actual brake torque balances the aerodynamic torque exactly (making
allowance for the gearbox ratio if the brake is on the high speed shaft) and there is no further
rotation. However, if the aerodynamic torque increases to overcome the maximum or applied
brake torque, the brake starts to slip and rotation recommences.
The rigid drive train model may be used in combination with flexible drive train mountings.
In this case the equations of motion are more complex - see Section 4.3.
4.1.3 Flexible shaft model
The flexible shaft model is obtained by selecting the dynamic drive train model with torsional
flexibility in one or both shafts. It allows separate degrees of freedom for the rotation of the
turbine rotor and the generator rotor. The torsional flexibility of the low speed and high
speed shafts may be specified independently. As with the rigid shaft model, a model of the
generator must be provided so that the generator reaction torque is specified.

19 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

The turbine rotor is accelerated by the torque imbalance between the aerodynamic torque
(adjusted for the effect of edgewise modes as explained in Section 4.1.2) and the low speed
shaft torque. The generator rotor is accelerated by the imbalance between high speed shaft
torque and generator reaction torque. The shaft torques are calculated from the shaft twist,
together with any applied brake torque contributions depending on the location of the brake,
which may be specified as being at either end of either the low or high speed shaft.
During a parked simulation, or once the brake disk has come to rest during a stopping
simulation, the equations of motion change depending on the brake location. If the brake is
immediately adjacent to the rotor or generator then there is no further rotation of that
component, but the other component continues to move and oscillates against the torsional
flexibility of the shafts. If the brake is adjacent to the gearbox and both shafts are flexible,
then both rotor and generator will oscillate. However, if the torque at the brake disk
increases to overcome the maximum or applied brake torque, then the brake starts to slip
again.
The flexible drive train model may be used in combination with flexible drive train
mountings. In this case the equations of motion are more complex - see Section 4.3.
It should be pointed out that while the flexible shaft model provides greater accuracy in the
prediction of loads, there is potential for one of the drive drain vibrational modes to be of
relatively high frequency, depending on the generator inertia and shaft stiffnesses. The
presence of this high frequency mode could result in slower simulations.

4.2 Generator models


The generator characteristics must be provided if either the rigid or flexible shaft drive train
model is specified. Three generator models are available:
A directly-connected induction generator model (for constant speed turbines),
A variable speed generator model (for variable speed turbines), and
A variable slip generator model (providing limited range variable speed above rated)
4.2.1 Fixed speed induction generator
This model represents an induction generator directly connected to the grid. Its
characteristics are defined by the slip slope h and the short-circuit transient time constant .
The air-gap or generator reaction torque Q is then defined by the following differential
equation:

Q& = 1 [h(
where

) Q]

is the actual generator speed and

is the generator synchronous or no-load speed.

The slip slope is calculated as

Pr

h=
r

20 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

where r is the generator speed at rated power output Pr , given by r =


S is the rated slip in %, and is the full load efficiency of the generator.

FINAL

(1 + S/100) where

4.2.2 Fixed speed induction generator: electrical model


A more complete model of the directly-connected induction generator is also available in
Bladed. This model requires the equivalent circuit parameters of the generator to be supplied
(at the operating temperature, rather than the cold values), along with the number of pole
pairs, the voltage and the network frequency. It is also possible to model power factor
correction capacitors and auxiliary loads such as turbine ancillary equipment. The equivalent
circuit configuration is shown in Figure 4.1.
Rr/s

Rs
xs

xr

xm

Ra
C
Xa

Rs = Stator resistance
xs = Stator reactance
Rr = Rotor resistance
xr = Rotor reactance
xm = Mutual reactance
C = Power factor correction
Ra = Auxiliary load resistance
Xa = Auxiliary load reactance
s = slip

Figure 4.1: Equivalent circuit model of induction generator


The equivalent circuit parameters should be given for a star-connected generator. If the
generator is delta-connected, the resistances and reactances should be divided by 3 to convert
to the equivalent star-connected configuration.
The voltage should be given as rms line volts. To convert peak voltage to rms, divide by 2.
To convert phase volts to line volts, multiply by 3.
Since this model necessarily includes electrical losses in the generator and ancillary
equipment, it is not possible to specify any additional electrical losses, although mechanical
losses may be specified - see Section 4.4.
Four different models of the electrical dynamics of the system illustrated in Figure 4.1 are
provided:

Steady state
1st order
2nd order
4th order

The steady state model simply calculates the steady-state currents and voltages in Figure 4.1
at each instant. The 1st order model introduces a first order lag into the relationship between
the slip (s) and the effective rotor resistance (Rr/s), using the short-circuit transient time
constant given by [4.1]:

X s X r x 2m
XsR r s

where Xs = xs+xm, Xr = xr+xm, and

is the grid frequency in rad/s.

21 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

The 2nd order model represents the generator as a voltage source


reactance X = Xs - xm2/Xr, ignoring stator flux transients:

FINAL

behind a transient

is (rs + jX) = vs where is and vs are the stator current and terminal voltage respectively. The dynamics of the
rotor flux linkage r may be written as

1
& r = rr i r + js
(
+
s
)
1
s

where s is the fractional slip speed (positive for generating) and ir is the rotor current. This
can be re-written in terms of the induced voltage using
xm
r = j
Xr
to give

T0 & =

rs + jX s
rs + jX

js

T0

+j

Xs X
vs
rs + jX

where
T0 =

Xr
.
s rr

The 4th order model is a full d-q (direct and quadrature) axis representation of the generator
which uses Parks transformation [4.2] to model the 3-phase windings of the generator as an
equivalent set of two windings in quadrature [4.3]. Using complex notation to represent the
direct and quadrature components of currents and voltages as the real and imaginary parts of
a single complex quantity, we can obtain

xsx r
s

x 2m d i s
dt i r

x r rs + jx 2m (1 + s)
x m rs

jx m x s (1 + s)

x m rr + jx m xr (1 + s)
x s rr

jx s x r (1 + s)

is
ir

xr
v
xm s

where all the currents and voltages are now complex.


Where speed of simulation is more important than accuracy, one of the lower order models
should be used. The 4th order model should be used for the greatest accuracy, although in
many circumstances the lower order models give very similar results. The lower order
models do not give an accurate representation of start-up transients, however.
4.2.3 Variable speed generator
This model should be used for a variable speed turbine incorporating a frequency converter to
decouple the generator speed from the grid frequency. The variable speed drive, consisting
of both the generator and frequency converter, is modelled as a whole. A modern variable
speed drive is capable of accepting a torque demand and responding to this within a very
short time to give the desired torque at the generator air-gap, irrespective of the generator
speed (as long as it is within specified limits). A first order lag model is provided for this
response:

22 of 82

Garrad Hassan and Partners Ltd

Qg =

Document: 282/BR/009

ISSUE:011

FINAL

Qd
(1 + e s)

where Qd is the demanded torque, Qg is the air-gap torque, and e is the time constant of the
first order lag. Note that the use of a small time constant may result in slower simulations. If
the time constant is very small, specifying a zero time constant will speed up the simulations,
without much effect on accuracy.
A variable speed turbine requires a controller to generate an appropriate torque demand, such
that the turbine speed is regulated appropriately. Details of the control models which are
available with Bladed can be found in Section 5.
The minimum and maximum generator torque must be specified. Motoring may occur if a
negative minimum torque is specified.
The phase angle between current and voltage, and hence the power factor, is specified, on the
assumption that, in effect, both active and reactive power flows into the network are being
controlled with the same time constant as the torque, and that the frequency converter
controller is programmed to maintain constant power factor.
An option for drive train damping feedback is provided. This represents additional
functionality which may be available in the frequency converter controller which adds a term
derived from measured generator speed onto the incoming torque demand. This term is
defined as a transfer function acting on the measured speed. The transfer function is supplied
as a ratio of polynomials in the Laplace operator, s. Thus the equation for the air-gap torque
Qg becomes

Qg =

Qd
Num(s)
+
(1 + e s) Den(s)

where Num(s) and Den(s) are polynomials. The transfer function would normally be some
kind of tuned bandpass filter designed to provide some damping for drive train torsional
vibrations, which in the case of variable speed operation may otherwise be very lightly
damped, sometimes causing severe gearbox loads.
4.2.4 Variable slip generator
A variable slip generator is essentially an induction generator with a variable resistance in
series with the rotor circuit [4.3, 4.4]. Below rated power, it acts just like a fixed speed
induction generator, so the same parameters are required as described in Section 4.2.1.
Above rated, the variable slip generator uses a fast-switching controller to regulate the rotor
current, and hence the air-gap torque, so the generator actually behaves just like a variable
speed system, albeit with a limited speed range. The same parameters as for a variable speed
system must therefore also be supplied (see Section 4.2.3), with the exception of the phase
angle since power factor control is not available in this case.
Alternatively, a full electrical model of the variable slip generator is available. The generator
is modelled as in Section 4.2.2, and the rotor current controller is modelled as a continuoustime PI controller which adjusts the rotor resistance between the defined limits (with

23 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

integrator desaturation on the limits), in response to the difference between the actual and
demanded rotor current. The steady-state relationship between torque and rotor current is
computed at the start of the simulation, so that the torque demand can be converted to a rotor
current demand. The scheme is shown in Figure 4.2.
Torque
demand

Current
demand

1
|I|

PI with
limits

Rotor
resistanc
e

Measured current |I|


Figure 4.2: Variable slip generator rotor current controller

4.3 Drive train mounting


If desired, torsional flexibility may be specified either in the gearbox mounting or between
the pallet or bedplate and the tower top. This option is only allowed if either the stiff or
flexible drive train model is specified, and it adds an additional rotational degree of freedom.
In either case, the torsional stiffness and damping of the mounting is specified, with the axis
of rotation assumed to coincide with the rotor shaft. The moment of inertia of the moving
components about the low speed shaft axis must also be specified. In the case of a flexible
gearbox mounting, this is the moment of inertia of the gearbox casing. In the case of a
flexible pallet mounting, it is the moment of inertia of the gearbox casing, the generator
stator, the moving pallet and any other components rigidly fixed to it.
If either form of mounting is specified, the direction of rotation of the generator shaft will
affect some of the internal drive train loads. If the low speed and high speed shafts rotate in
opposite directions, specify a negative gearbox ratio in the drive train model. The effect of
any offset between the low speed shaft and high speed shaft axes is ignored.
Any shaft brake is assumed to be rigidly mounted on the pallet. Thus any motion once the
brake disk has stopped turning depends on the type of drive train mounting as well as on the
position of the brake on the low or high speed shaft. For example if there is a soft pallet
mounting, then there will still be some oscillation of the rotor after the brake disk has stopped
even if both shafts are stiff.
As in the case of the flexible shaft drive train model, it should be pointed out that while
modelling the effect of flexible mountings provides greater accuracy in the prediction of
loads, there is potential for one or two of the resulting drive train vibrational modes to be of
relatively high frequency, depending on the various moments of inertia and shaft and
mounting stiffnesses. The presence of high frequency modes could result in slower
simulations.

4.4 Energy losses


Power train energy losses are modelled as a combination of mechanical losses and electrical
losses in the generator (including the frequency converter in the case of variable speed
turbines).

24 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

Mechanical losses in the gearbox and/or shaft bearings are modelled as either a loss torque or
a power loss, which may be constant, or interpolated linearly from a look-up table. This may
be a look-up table against rotor speed, gearbox torque or shaft power, or a two-dimensional
look-up table against rotor speed and either shaft torque or power. Mechanical losses
modelled in terms of power are inappropriate if calculations are to be carried out at low or
zero rotational speeds, e.g. for starts, stops, idling and parked calculations. In these cases, the
losses are better expressed in terms of torque.
The electrical losses may specified by one of two methods:
Linear model: This requires a no-load loss LN and an efficiency , where the electrical power
output Pe is related to the generator shaft input power Ps by:
Pe =

(Ps - LN)

Look-up table: The power loss L(Ps) is specified as a function of generator shaft input power
Ps by means of a look-up table. The electrical power output Pe is given by:
Pe = Ps - L(Ps)
Linear interpolation is used between points on the look-up table.
Note that if a full electrical model of the generator is used, additional electrical losses in this
form cannot be specified since the generator model implicitly includes all electrical losses.

4.5 The electrical network


Provided either the detailed electrical model of the induction generator or the variable speed
generator model is used, so that electrical currents and voltages are calculated, and reactive
power as well as active power, then the characteristics of the network to which the turbine is
connected may also be supplied. As well as allowing the voltage variations, and hence the
flicker, at various points on the network to be calculated, the presence of the network may
also, in the case of the directly connected induction generator, influence the dynamic
response of the generator itself particularly on a weak network.
The network is modelled as a connection, with defined impedance, to the point of common
coupling (PCC in Figure 4.2) and a further connection, also with defined impedance, to an
infinite busbar. Further turbines may be connected at the point of common coupling. These
additional turbines are each assumed to be identical to the turbine being modelled, including
the impedance of the connection to the point of common coupling. However they are
modelled as static rather dynamic, with current and phase angle constant during the
simulation. The initial conditions are calculated with the assumption that all turbines are in
an identical state, and the other turbines then remain in the same state throughout. Thus the
steady state voltage rise due to all the turbines at the point of common coupling will be taken
into account in calculating the performance of the turbine whose performance is being
simulated .

25 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

Other turbines
(if required)
Wind
turbine

R1 + jX1
Windfarm
interconnection
impedance

PCC

Figure 4.2: The network model

26 of 82

R2 + jX2
Network
connection
impedance

Infinite
busbar

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

5. CLOSED LOOP CONTROL

5.1 Introduction
Closed loop control may be used during normal running of the turbine to control the blade
pitch angle and, for variable speed turbines, the rotor speed. Four different controller types
are provided:
1. Fixed speed stall regulated. The generator is directly connected to a constant frequency
grid, and there is no active aerodynamic control during normal power production.
2. Fixed speed pitch regulated. The generator is directly connected to a constant frequency
grid, and pitch control is used to regulate power in high winds.
3. Variable speed stall regulated. A frequency converter decouples the generator from the
grid, allowing the rotor speed to be varied by controlling the generator reaction torque. In
high winds, this speed control capability is used to slow the rotor down until aerodynamic
stall limits the power to the desired level.
4. Variable speed pitch regulated. A frequency converter decouples the generator from the
grid, allowing the rotor speed to be varied by controlling the generator reaction torque. In
high winds, the torque is held at the rated level and pitch control is used to regulate the
rotor speed and hence also the power.
For a constant speed stall regulated turbine no parameters need be defined as there is no
control action. In the other cases the control action will determine the steady state operating
point of the turbine as well as its dynamic response. For steady state calculations it is only
necessary to specify those parameters which define the operating curve of the turbine. For
dynamic calculations, further parameters are used to define the dynamics of the closed loop
control. The parameters required are defined further in the following sections.
Note that all closed loop control data are defined relative to the high speed shaft.

5.2 The fixed speed pitch regulated controller


This controller is applicable to a turbine with a directly-connected generator which uses blade
pitch control to regulate power in high winds. It is applicable to full or partial span pitch
control, as well as to other forms of aerodynamic control such as flaps or ailerons. In the
latter case, the pitch angle can be taken to refer to the deployment angle of the flap or aileron.
From the optimum position, the blades may pitch in either direction to reduce the
aerodynamic torque. If feathering pitch action is selected, the pitchable part of the blade
moves to reduce its angle of attack as the wind speed (and hence the power) increases. If
stalling pitch action is selected, it moves in the opposite direction to stall the blade as the
wind speed increases. In the feathering case, the minimum pitch angle defines the pitch
setting below rated, while in the stalling case the maximum pitch angle is used below rated,
and the pitch decreases towards the minimum value (usually a negative pitch angle) above
rated.

27 of 82

Garrad Hassan and Partners Ltd

Wind

Document: 282/BR/009

Electric
power

Turbine

Blade pitch

ISSUE:011

FINAL

Power
Measured
transducer power Controlle

Pitch
actuator

Pitch
demand

Power
set-point

Figure 5.1: The fixed speed pitch regulated control loop

Figure 5.1 shows schematically the elements of the fixed speed pitch regulated control loop
which are modelled.
5.2.1 Steady state parameters
In order to define the steady-state operating curve, it is necessary to define the power setpoint and the minimum and maximum pitch angle settings, as well as the direction of pitching
as described above. The correct pitch angle can then be calculated in order to achieve the setpoint power at any given steady wind speed.
5.2.2 Dynamic parameters
To calculate the dynamic behaviour of the control loop, it is necessary to specify the dynamic
response of the power transducer and the pitch actuator, as well as the actual algorithm used
by the controller to calculate a pitch demand in response to the measured power signal.
Section 5.5 describes the available transducer and actuator models, while Section 5.6
describes the PI algorithm which is used by the controller.

5.3 The variable speed stall regulated controller


This controller model is appropriate to variable speed turbines which employ a frequency
converter to decouple the generator speed from the fixed frequency of the grid, and which do
not use pitch control to limit the power above rated wind speed. Instead, the generator
reaction torque is controlled so as to slow the rotor down into stall in high wind speeds. The
control loop is shown schematically in Figure 5.2.
5.3.1 Steady state parameters
The steady-state operating curve can be described with reference to a torque-speed graph as
in Figure 5.3. The allowable speed range in the steady state is from S1 to S2. In low winds it
is possible to maximise energy capture by following a constant tip speed ratio load line which
corresponds to operation at the maximum power coefficient. This load line is a quadratic
curve on the torque-speed plane, shown by the line BG in Figure 5.3. Alternatively a look-up
table may be specified. If there is a minimum allowed operating speed S1, then it is no
longer possible to follow this curve in very low winds, and the turbine is then operated at
nominally constant speed along the line AB shown in the figure. Similarly in high wind
speeds, once the maximum operating speed S4 is reached, then once again it is necessary to
28 of 82

Garrad Hassan and Partners Ltd

Wind

Document: 282/BR/009

ISSUE:011

Generator
speed

Speed
transducer

Electrical
power

Power
Measured Controlle
transducer
power

FINAL

Measured
speed

Turbine

Generator
torque
demand

Desired
power,
torque,
speed

Figure 5.2: The variable speed stall regulated control loop


depart from the optimum load line by operating at nominally constant speed along the line
GH.
Once maximum power is reached at point H, it is necessary to slow the rotor speed down into
stall, along the constant power line HI. If high rotational speeds are allowed, it is of course
possible for the line GH to collapse so that the constant power line and the constant tip speed

Figure 5.3: Variable speed stall regulated operating curve

29 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

ratio line meet at point J.


Clearly the parameters needed to specify the steady state operating curve are:
The minimum speed, S1
The maximum speed in constant tip speed ratio mode, S4
The maximum steady-state operating speed. This is usually S4, but could conceivably be
higher in the case of a turbine whose characteristics are such that as the wind speed
increases, the above rated operating point moves from H to I, then drops back to H, and
then carries on (towards J) in very high winds. This situation is somewhat unlikely
however, because if rotational speeds beyond S4 are permitted in very high winds, there is
little reason not to increase S4 and allow the same high rotor speeds in lower winds.)
The above rated power set-point, corresponding to the line HI. This is defined in terms of
shaft power. Electrical power will of course be lower if electrical losses are modelled.
The parameter K which defines the constant tip speed ratio line BG. This is given by:
K =

R5 Cp( ) / 2

G3

where
= air density
R = rotor radius
= desired tip speed ratio
Cp( ) = Power coefficient at tip speed ratio
G = gearbox ratio
Then when the generator torque demand is set to K 2 where is the measured generator
speed, this ensures that in the steady state the turbine will maintain tip speed ratio and the
corresponding power coefficient Cp( ). Note that power train losses may vary with rotational
speed, in which case the optimum rotor speed is not necessarily that which results in the
maximum aerodynamic power coefficient.
As an alternative to the parameter K , a look-up table may be specified giving generator
torque as a function of speed.
5.3.2 Dynamic parameters
To calculate the dynamic behaviour of the control loop, it is necessary to specify the dynamic
response of both power and speed transducers, as well as the actual algorithm used by the
controller to calculate a generator torque demand in response to the measured power and
speed signals. Section 5.5 describes the available transducer and actuator models.
Two closed loop control loops are used for the generator torque control, as shown in Figure
5.4. An inner control loop calculates a generator torque demand as a function of generator
speed error, while an outer loop calculates a generator speed demand as a function of power
error. Both control loops use PI controllers, as described in Section 5.6.
Below rated, the speed set-point switches between S1 and S4. In low winds it is at S1, and
the torque demand output is limited to a maximum value given by the optimal tip speed ratio
curve BG. This causes the operating point to track the trajectory ABG. In higher winds, the
set-point changes to S4, and the torque demand output is limited to a minimum value given by
the optimal tip speed ratio curve, causing the operating point to track the trajectory BGH.
Once the torque reaches QR, the outer control loop causes the speed set-point to reduce along
HI, and the inner loop tracks this varying speed demand.
30 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

PI
controller

Power
set-point

ISSUE:011

FINAL

PI
controller

Speed
demand

Measured power
Measured speed
Generator torque demand
Figure 5.4: Stall regulated variable speed control loops

5.4 The variable speed pitch regulated controller


This controller model is appropriate to variable speed turbines, which employ a frequency
converter to decouple the generator speed from the fixed frequency of the grid, and which use
pitch control to limit the power above rated wind speed. The control loop is shown
schematically in Figure 5.5.
5.4.1 Steady state parameters
The steady-state operating curve can be described with reference to the torque-speed graph
shown in Figure 5.6. Below rated, i.e. from point A to point H, the operating curve is exactly
as in the stall regulated variable speed case described in Section 5.3.1, Figure 5.3. Above
rated however, the blade pitch is adjusted to maintain the chosen operating point, designated

Wind

Generator
speed

Speed
transducer

Measured
speed

Turbine
Controlle
Blade pitch

Pitch
actuator

Pitch
demand

Generator
torque
demand

Desired
torque
and speed

Figure 5.5: The variable speed pitch regulated control loop


L. Effectively, changing the pitch alters the lines of constant wind speed, forcing them to
pass through the desired operating point.

31 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

Figure 5.6: Variable speed pitch regulated operating curve

Once rated torque is reached at point H, the torque demand is kept constant for all higher
wind speeds, and pitch control regulates the rotor speed. A small (optional) margin is
allowed between points H (where the torque reaches maximum) and L (where pitch control
begins) to prevent excessive mode switching between below and above rated control modes.
However, this margin may not be required, in which case points H and L coincide. As with
the stall regulated controller, the line GH may collapse to a point if desired.
Clearly the parameters needed to specify the steady state operating curve are:
The minimum speed, S1
The maximum speed in constant tip speed ratio mode, S4
The speed set-point above rated (S5). This may be the same as S4.
The maximum steady-state operating speed. This is normally the same as S5.
The above rated torque set-point, QR.
The parameter K which defines the constant tip speed ratio line BG, or a look-up table.
This is as defined in Section 5.3.1.
5.4.2 Dynamic parameters
To calculate the dynamic behaviour of the control loop, it is necessary to specify the dynamic
response of the speed transducer and the pitch actuator, as well as the actual algorithm used
by the controller to calculate the pitch and generator torque demands in response to the
measured speed signal. Section 5.5 describes the available transducer and actuator models.

32 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

Figure 5.7 shows the control loops used to generate pitch and torque demands. The torque
demand loop is active below rated, and the pitch demand loop above rated. Section 5.6
describes the PI algorithm which is used by both loops.
Below rated, the speed set-point switches between S1 and S4. In low winds it is at S1, and
the torque demand output is limited to a maximum value given by the optimal tip speed ratio
curve BG. This causes the operating point to track the trajectory ABG. In higher winds, the
set-point changes to S4, and the torque demand output is limited to a minimum value given by
the optimal tip speed ratio curve, causing the operating point to track the trajectory BGH, and
a maximum value of QR. When point H is reached the torque remains constant, with the
pitch control loop becoming active when the speed exceeds S5.
Above rated
Speed
set-point
Below rated

Measured speed
Blade

PI
controller

PI
controller

pitch

Generator torque demand


Figure 5.7: Pitch regulated variable speed control loops

5.5 Transducer models


First order lag models are provided in Bladed to represent the dynamics of the power
transducer and the generator speed transducer. The first order lag model is represented by
y& =

1
(x
T

y)

where x is the input and y is the output. The input is the actual power or speed and the output
is the measured power or speed, as input to the controller.

5.6 Modelling the pitch actuator


The pitch actuator may be modelled as either a pitch position or pitch rate actuator, and either
active or passive dynamics may be specified.
The simplest model is a passive actuator, with the relationship between the input and the
output represented by a transfer function. For the pitch position actuator, the input is the
pitch demand generated by the controller and the output is the actual pitch angle of the
blades. For the pitch rate actuator, the input is the pitch rate demand generated by the
controller and the output is the actual pitch rate at which the blades move. The transfer

33 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

function may be a first order lag, a second order response, or a general transfer function, up
to 8th order.
The first order lag model is represented by
y& =

1
(x
T

y)

where x is the input and y is the output. The second-order model is represented by
&&
y + 2! y& =

(x

y)

where is the bandwidth and ! the damping factor. The general transfer function model is
represented by numerator and denominator polynomials in the Laplace operator.
For detailed calculations, especially to understand the loads on the pitch actuator itself and
the duty which will be required of it, it is possible to enter a more detailed model. This can
take into account any internal closed loop dynamics in the actuator, and also the pitch motion
resulting from the actuator torque acting on the pitching inertia, with or against the
aerodynamic pitch moment and the pitch bearing friction. The bearing friction itself depends
critically on the loading at the pitch bearing.
Figure 5.8 shows the various options for controlling the pitch angle, starting from either a
pitch position demand or a pitch rate demand. The pitch position demand may optionally be
processed through a ramp control, shown in Figure 5.9, which smooths the step changes in
demand generated by a discrete controller by applying rate and/or acceleration limits. Then
the pitch position demand can act either through passive dynamics to generate a pitch
position, or through a PID controller on pitch error to generate a pitch rate demand. Rate
limits are applied to the output, with instantaneous integrator desaturation to prevent wind-up
in the PID case. Thus the pitch rate demand may come either from here or directly from the
controller. This rate demand can act either through passive dynamics to generate a pitch rate,
or through a PID controller on pitch rate error to generate an actuator torque demand. In the
latter case, the pitch actuator passive dynamics then generate an actual actuator torque, which
acts against bearing friction and any aerodynamic pitching moment to accelerate the pitching
inertia of the blades and the actuator itself. An optional first order filter on each PID input
allows step changes in demand from the controller to be smoothed, and instantaneous
integrator desaturation prevents wind-up when the torque limits are reached.
Both PID controllers include a filter on the differential term to prevent excessive high
frequency gain. Also there is a choice of derivative action, such that the derivative gain may
be applied either to the feedback (i.e. the measured position or rate), the error signal, or the
demand. The latter case represents a feed-forward term in the controller.
If passive pitch rate dynamics are selected, the response will be subject to acceleration limits
calculated from the aerodynamic pitching moment, bearing friction and the actuator toque
limits acting on the pitching inertia. If the total pitching inertia is zero, no limits will be
applied.
The pitch bearing sliding friction torque is modelled as the sum of four terms: a constant, a
term proportional to the bending moment at the bearing, and a terms proportional to the axial
and radial forces on the bearing. Sometimes the actuator cannot overcome the applied

34 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

torques and the pitch motion will stick. Before it can move again, the break-out or stiction
torque must be overcome. This is modelled as an additional contribution to the friction
torque while the pitch is not moving. This additional contribution is specified as a constant
torque, plus a term proportional to the sliding friction torque.

Pitch position
demand from
controller

Measured
pitch
position

Pitch rate
demand from
controller

Measured
pitch rate

Bearing
loads

Ramp control

Pitching
moment
PID controller

Actuator
torque
limits
Pitch rate
demand
+
Pitching
inertia
Acceleration
limits

Passive
dynamics

PID controller

Actuator
torque
demand

Passive
dynamics

Passive
dynamics
Actuator
torque
Pitching
inertia

Pitch rate

Actual pitch
position

Figure 5.8: Pitch actuator options

35 of 82

Bearing
friction

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

1.2

Demand

0.8

0.6

0.4

Raw demand
Rate limit

0.2

Acceleration limit
Rate & acceleration limits

0
-0.2

0.2

0.4

0.6

0.8

1.2

Timesteps

The ramp is re-started each timestep. If the ramp is not completed by the end of the timestep
and an acceleration limit is specified, the slope at the start of the next timestep will be nonzero.
Figure 5.9: Ramp control for pitch actuator position demand

5.7 The PI control algorithm


All the closed loop control algorithms described above use PI controllers to calculate the
output y (pitch, torque or speed demand) from the input x (power or speed error). The basic
PI algorithm can be expressed as

y& = K p x& + Ki x
where Kp and Ki represent the proportional and integral gains. The ratio Kp/Ki is also known
as the integral time constant. Calculation of appropriate values for the gains is a specialist
task, which should take into account the dynamics of the wind turbine together with the
aerodynamic characteristics and principal forcing frequencies, and should aim to achieve
stable control at all operating points and a suitable trade-off between accuracy of tracking the
set-point and the degree of actuator activity.
Straightforward implementation of the above equation leads to the problem of integrator
wind-up if the output y is subject to limits, as is the case here. This means that the raw
output calculated as above continues to change as a result of the integral (Ki) term even
though the actual output is being constrained to a limit. When the direction of movement of y
changes, it will then take a long time before it comes back to the limit so that the final
(constrained) output starts to change. This is avoided in the continuous-time implementation
of the PI controller by an additional term -"y/Td in the above equation, where "y is the

36 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

amount by which the raw output y has gone beyond the limit, and Td is the desaturation time
constant which must be supplied by the user.
In practice the control algorithm is usually implemented in a digital controller working on a
discrete timestep. In the Bladed model, the continuous implementation of the controller is an
approximate representation, although the discrete timestep is usually fast enough for the
approximation to be a very good one. Since the integrator desaturation in a discrete
controller can be implemented by fully adjusting the raw integrator output at every timestep,
a suitable approximation for the continuous case is to use a desaturation time constant
approximately equal to the discrete controller timestep.
Alternatively, perfect or instantaneous desaturation can be specified by setting the
desaturation time constant to zero.
5.7.1 Gain scheduling
Since the characteristics of the turbine, especially the aerodynamic characteristics, are not
constant but will vary according to the operating point, and hence the wind speed, it may be
necessary to adjust the controller gains as a function of the operating point in order to ensure
that suitable control loop characteristics are achieved at all wind speeds. This is known as
gain scheduling, and the gain scheduling model provided in Bladed allows both the
proportional and integral gains of any control loop to be scaled by a factor 1/F, where F is a
function of some variable V which is accessible to the controller and which is representative
of the operating point in some way.
The choices available are:
F = constant
F = F(V) as defined by a look-up table
F = F(V) as defined by a polynomial, but with minimum and maximum limits applied to F
The choice of variable V depends on the particular control loop. The following choices are
provided:
Fixed speed pitch regulated controller:
Electrical power, pitch angle, wind speed.
Variable speed below-rated torque controller:
Electrical power, generator speed, wind speed, and pitch angle (in the pitch regulated
case).
Variable speed stall regulated above-rated controller:
Electrical power, generator speed, wind speed.
Variable speed pitch regulated above-rated controller:
Electrical power, generator speed, wind speed, pitch angle.
The variables shown in bold are normally recommended. Gain scheduling is unlikely to be
required for the variable speed below rated controllers. For the variable speed stall regulated
above-rated controller, no general rule can be given. Gain scheduling on wind speed is not
usually a practical proposition because of the difficulty of measuring a representative wind
speed, and this option is only provided for research purposes. The wind speed used is the

37 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

hub wind speed, which may differ from any wind speed measured by an anemometer
mounted on the nacelle, especially in the case of an upwind turbine.
Gain scheduling on pitch angle is recommended for the pitch regulation controllers, to
compensate for the large changes in the sensitivity of aerodynamic torque to pitch angle over
the operating range. The steady loads calculation may be used to calculate the partial
derivative of aerodynamic torque with respect to pitch angle, and F may be set proportional
to this. In many cases, simply setting F proportional to pitch angle is a good approximation,
but a lower limit for F must be set to prevent excessive gains at small pitch angles.

5.8 Control mode changes


The variable speed controllers, both stall regulated and pitch regulated, require the following
mode changes:
Change of speed set-point from S1 to S4 (refer to Figures 5.3 and 5.6). This occurs when
the measured speed crosses the threshold value (S1+S4)/2. This mode change is
completely benign as the control action along the optimum tip speed ratio line BG is the
same either side of the mode change point, so no hysteresis is required.
Change from below rated to above rated control.
For the stall regulated case, the change from below rated to above rated is also benign.
Making the switch in the middle of the section GH of Figure 5.3 causes no immediate change
in control action. However, in the case of G and H coinciding, or being very close together, it
may be necessary to modify the mode change strategy, depending on the turbine
characteristics.
For the pitch regulated case, the change to above rated control occurs when the torque
demand is at maximum (QR) and the speed exceeds S5 (refer to Figure 5.6). The change to
below rated occurs when the pitch demand is at fine pitch (minimum pitch for the feathering
case, maximum pitch for pitch-assisted stall) and the speed falls below S4. While this
strategy is usually suitable, it may be desirable to modify it depending on the turbine
characteristics.
The mode changes occur on a discrete timestep set to a default value of 0.1 seconds.

5.9 Client-specific controllers


The control algorithms described above have been developed to be suitable for a wide range
of cases. However, it is recognised that there is great variation in the design of controllers for
wind turbines. In a number of specific cases, Garrad Hassan have enhanced these basic
controller designs in various ways to suit particular turbine designs, further improving the
control performance. In many cases, particularly for variable speed turbines, both the closed
loop performance and the mode changing behaviour can be improved significantly with a
small additional degree of sophistication.

38 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

It is also recognised that simulations must be able to be adapted to use any particular
controller design, both to allow algorithms different from the standard ones described above,
and also to allow the modelling of discrete controllers, for example so that the effect of
controller timestep can be investigated.
For these reasons, Bladed offers the possibility of incorporating user-defined controllers in
the dynamic simulations. Through a defined interface which makes use of a shared file, a
users control program, written in any language, can be used to control the simulation.
The user-defined controller may do any of the following:
Blade pitch angle or blade pitch rate control during any phase of operation including
power production, stops, starts, idling etc.
Generator torque control for variable speed turbines
Control the generator contactor, allowing the generator to be switched on or off for
simulating stops and starts
Control the shaft brake, to simulate transitions between parked, idling, starting, stopping,
and power production states.
Control of nacelle yaw to simulate closed loop yaw control algorithms and/or yawing
strategies for start-up, shutdown etc.
The User Manual describes how to write a user-defined control program.

5.10 Signal noise and discretisation


When a discrete external controller is used, Bladed offers the possibility of adding random
noise to the measured signals sent to the controller, and also to discretise the signals to a
specified resolution.
The random noise may be Gaussian, in which case the standard deviation of the noise must
be specified, or it may be from a rectangular distribution, in which case the half-width of the
distribution should be given. The noise is added to the signal before it is discretised.

39 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

6. SUPERVISORY CONTROL
This section of the manual covers the modelling of the following aspects of turbine operation:

Start-up
Shut-down (normal and emergency stops)
Non-operational situations (rotor parked or idling)
Operation of the shaft brake
Teeter restraints
Yaw control

The standard implementation of these features in the simulation model is described. As in


the case of Closed Loop Control, alternative supervisory control logic can be incorporated in
a user-defined controller - see Section 5.9.

6.1 Start-up
Simulation of a wind turbine start-up begins with the rotor at a specified speed (usually but
not necessarily zero) and the generator off-line. The brake is assumed to be released at the
start of the simulation (i.e. at time zero).
If blade pitch or aileron control is available, the initial pitch or aileron angle is specified,
along with a constant rate of change which continues until either a specified angle is reached
or the closed loop controller takes over.
When a specified rotational speed is reached, the generator comes on line, and the closed
loop controller begins to operate. The simulation continues until the specified simulation end
time.
In the case of a variable speed turbine, there may be a transition period after cut-in of the
closed loop controller before the turbine is fully in the normal running state. There are two
different cases:
Variable speed pitch regulation: in the case when the pitch angle has not yet reached the
normal operating value (fine pitch) at the moment when the closed loop controller cuts in,
then the pitch change rate for start-up continues to apply until either fine pitch is reached, or
until the conditions of Section 5.8 for starting the closed loop pitch controller are satisfied.
Variable speed stall regulation: when the closed loop controller cuts in, the above-rated
control mode is assumed to apply initially. In practice this assumption does not affect the
start-up since in low winds the operating point would be constrained by the quadratic
optimum-Cp characteristic in any case.

40 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

6.2 Normal stops


A normal stop is initiated at a specified time after the start of the simulation. Normal
operation in power production mode is assumed prior to this point, with full structural and
control dynamics in effect if desired. The structural dynamics continue in effect during the
entire simulation.
The standard logic for a normal stop is to start pitching the blades (or moving the ailerons) at
a specified rate from the moment that the stop is initiated, continuing until a final pitch angle
is reached. The generator is taken off-line when the electrical power reaches zero in the case
of a fixed-speed turbine, or when the minimum generator speed is reached in the case of a
variable speed turbine.
Once the rotational speed drops below a specified value, the shaft brake is applied to bring
the rotor to rest.
The simulation continues until the rotor comes to rest, or for a certain time longer if so
desired in order that the transient loads can be simulated as the brake disk stops. However,
the simulation end time overrides this, so it must be set long enough for the stop event to be
completed.
If there is no pitch control, the brake trip speed may be set high so that the shaft brake is
applied immediately at the initiation of the stop.
Section 6.4 describes the dynamic characteristics of the shaft brake itself.

6.3 Emergency stops


An emergency stop is initiated at a specified time after the start of the simulation. Normal
operation in power production mode is assumed prior to this point, with full structural and
control dynamics in effect if so desired. The structural dynamics continue in effect during
the entire simulation.
Several options are available for simulating emergency stops. In all cases it is assumed that
the generator load is lost at the initiation of the emergency stop, whether because of grid
failure or some electrical or mechanical failure of the turbine.
Pitch (or aileron) action is initiated either immediately or when the rotational speed exceeds a
specified value. A fixed pitch rate then applies until a final pitch angle is reached. Provision
is made for the pitch of one or more of the blades to stick at a specified angle to simulate
failure of a pitch bearing or actuator.
The shaft brake can also be applied either at the initiation of the stop or when a specified
overspeed is reached. Section 6.4 describes the dynamic characteristics of the shaft brake
itself. There is also a rotational speed below which the shaft brake is applied for parking, in
the event that it has not already been applied because of load loss or overspeed.
The simulation continues until the rotor comes to rest, or for a certain time longer if so
desired in order that the transient loads can be simulated as the brake disk stops. However,

41 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

the simulation end time overrides this, so it must be set long enough for the stop event to be
completed.

6.4 Brake dynamics


When the shaft brake is applied, either during a normal or an emergency stop, the full braking
torque is not available instantly. Instead, the torque builds up to the full value over a short
period of time. This torque build-up may be modelled as either a linear torque ramp, or by
specifying a look-up table giving achieved braking torque as a function of time.

6.5 Idling and parked simulations


For simulations in the idling and parked states, a fixed pitch angle is specified, the generator
is off line, and there is no pitch control action. In the case of a parked rotor the shaft brake is
applied, and the rotor azimuth must be specified. The azimuth is measured from zero with
blade 1 at top dead centre.
All specified structural dynamics will be in effect during these simulations. This also allows
for the possibility of the shaft brake slipping during a parked simulation if the shaft torque
exceeds the specified brake torque.

6.6 Yaw control


6.6.1 Active yaw
Active yaw movement may be specified in one of two ways:
1. One fixed-rate yaw manoeuvre may be specified, starting at a given point in any
simulation. This represents a change in the nominal nacelle position through a given
angle at a specified angular speed.
2. A user-defined controller (Section 5.9) may be used to specify either the yaw rate or the
yaw actuator torque at any time.
If active yaw is used to control the yaw rate, the effect of this is to change the demanded
nacelle angle in a specified way. The actual nacelle angle depends on the yaw dynamics see next section.

42 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

6.6.2 Yaw dynamics


Three options are available to define the yaw dynamics:
1. Rigid yaw: the actual nacelle angle exactly follows the demanded nacelle angle 0.
2. Flexible yaw: a certain amount of flexibility is present, usually in the yaw actuation
system, such that the actual nacelle angle may not follow the demanded nacelle angle
0 exactly. The extreme case is free yaw, when the demanded nacelle angle does not have
any effect.
3. Controlled yaw torque: this is available only with an external controller to define the yaw
actuator torque demand

Demanded yaw rate


Aerodynamic and
inertial yaw torque

Tower
Controlled
torque

Yaw spring

Friction

Yaw control type


None
Rigid
Flexible
Controlled torque

Demanded
yaw rate
No
Yes
Yes
No

Damper

Yaw spring and


damper
No
No
Yes
No

Friction
No
No
Yes
Yes

Controlled
torque
No
No
No
Yes

In the case of flexible or free yaw, the yaw damping Dy may be specified. This specifies a
torque Qd which opposes the yaw motion, given by

Qd = Dy ( & 0

&)

In the case of flexible yaw, a yaw spring may be specified either as a linear spring or as a
hydraulic accumulator system such as is often used to provide flexibility in hydraulic yaw
drives. The hydraulic system is assumed to be double-acting, with one accumulator (or set of
accumulators) on either side of the yaw motor. The torque opposing the motion is provided
by compression of the gas in the accumulators. If the nominal gas volume is V0 and the
instantaneous gas volumes either side of the yaw motor are v1 and v2 then the opposing torque
Qk is given by

Q k = KP0

V0
v1

V0
v2

43 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

where v1 = V0 + F( - 0 ) and v2 = V0 - F( - 0 ) and P0 is the equilibrium pressure in the


hydraulic system. The constant K defines the relationship between the torque developed at
the yaw bearing and the pressure difference across the yaw motor, while F the relationship
between the volume of oil flowing through the yaw motor and the resulting angular
movement at the yaw bearing. # is the gas law constant: PV# = RT. Putting # = 1 specifies
isothermal conditions in the accumulators.

6.7 Teeter restraint


Although not strictly a supervisory control function, the teeter restraint model available in
Bladed for teetered rotors is described here. The model allows a linear variation of restoring
torque with teeter angle, but also allows a free teeter range and an initial pre-load. Figure 6.1
defines the relevant parameters. Linear damping is also allowed, giving an additional torque
contribution proportional to teeter rate.
Restoring torque

Pre-load

Spring constant

free
teeter
angle

Figure 6.1: Teeter restraint model

44 of 82

Teeter angle

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

7. MODELLING THE WIND


The wind field incident on the turbine may be specified in a number of ways. For some
simple calculations, a uniform, constant wind speed is assumed, such that the same incident
wind speed is seen by every point on the rotor. For more detailed calculations however, it is
important to be able to define both the spatial and temporal variations in wind speed and
direction.
The steady-state spatial characteristics of the wind field may include any combination of the
following elements:
Wind shear: the variation of wind speed with height.
Tower shadow: distortion of the wind flow by the wind turbine tower.
Upwind turbine wake: full or partial immersion of the turbine rotor in the wake of
another turbine operating further upwind.
The wind direction must also be specified, both relative to the direction in which the nacelle
is pointing (to define the yaw error), and relative to the horizontal plane (to define the upflow
angle). The latter effect may be important for turbines operating in hilly terrain.
For simulations, it is also important to be able to define how the wind speed and direction
vary with time. The following alternative models are provided:
Constant wind: no variation with time.
Single point history: a time history of wind speed and direction, which is fully coherent
over the whole rotor, is specified as a look-up table against time. Linear interpolation is
used between the time points.
3D turbulent wind: this option uses a 3-dimensional turbulent wind field with defined
spectral and spatial coherence characteristics representative of real atmospheric
turbulence. This option will give the most realistic predictions of loads and performance
in normal conditions.
IEC transients: this option uses wind speed and direction transients as defined by the IEC
1400-1 standard [7.1, 7.7]. It is intended for evaluating specific load cases, for example
during extreme gusts.

45 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

7.1 Wind shear


Wind shear is the variation of steady state mean wind speed with height. Two alternative
models are provided, to relate the wind speed V(h) at height h above the ground to the wind
speed V(h0) at some reference height h0..
7.1.1 Exponential model
This model is defined in terms of a wind shear exponent $:

h
V (h) = V (h0 )
h0

Specifying the exponent as zero results in no wind speed variation with height.
7.1.2 Logarithmic model
This model is defined in terms of the ground roughness length z0:

V (h) = V (h0 )

log(h / z0 )
log(h0 / z0 )

7.2 Tower shadow


Tower shadow defines the distortion of the steady-state mean wind field due to the presence
of the wind turbine tower. Three different models are available: a potential flow model for
upwind rotors, an empirical tower wake model for downwind rotors, and a combined model
which is useful if the rotor yaws in and out of the downwind shadow area.
7.2.1 Potential flow model
This model is appropriate for rotors operating upwind of the tower. The longitudinal wind
velocity component upwind of the tower (V0) is modified using the assumption of
incompressible laminar flow around a cylinder of diameter D = F.DT where DT is the tower
diameter at the height where the tower shadow is being calculated, and F is a tower diameter
correction factor supplied by the user. For a point at a distance z in front of the tower
centreline and x to the side of the wind vector passing through the centreline, the wind speed
V is given by:

V ( x , z ) = AV0
where

A = 1+

D
2

( x2 z2 )
( x2 + z2 )2

provided the point is at an azimuth within +60 from bottom dead centre relative to the hub
centre. For azimuth within +60 of top dead centre it is assumed that V(x,z) = V0 , and to
46 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

ensure a smooth transition between these two zones, for all other azimuths, the factor A is
modified to A( 0.5 cos( )) + (0.5 + cos( )) where is the blade azimuthal position
7.2.2 Empirical model
For rotors operating downwind of the tower, an empirical model is provided, based on the
work of Powles [7.2] which uses a cosine bell-shaped tower wake. For a point at a distance z
behind the tower centreline and x to the side of the wind vector passing through the
centreline, the wind speed V is given by:

V ( x , z ) = AV0
where
A = 1 " cos 2

x
WDT

for azimuth angles within +60 of bottom dead centre. For other azimuth angles, the same
correction is applied as for the potential flow model, Section 7.2.2. Here " is the maximum
velocity deficit at the centre of the wake as a fraction of the local wind speed, and W is the
width of the tower shadow as a proportion of the local tower diameter DT. These quantities
are defined for a given downwind distance, also expressed as a proportion of DT . For other
distances, W increases, and " decreases, with the square root of the distance.
7.2.3 Combined model
The combined model simply uses the potential flow model at the front and sides of the tower,
and whichever of the other models gives the larger deficit at any point downwind. To ensure
a smooth transition, the product of the A factors of the two models is used in any small areas
where the potential flow model gives accelerated flow and the empirical model gives a
velocity deficit.

7.3 Upwind turbine wake


If the turbine rotor being modelled is assumed to be wholly or partially immersed in the wake
of another turbine operating further upwind, a model is provided to define the modification to
the steady-state mean wind profile caused by that wake.
A Gaussian profile is used to describe the wake of the upstream turbine. The local velocity at
a distance r from the wake centreline (which may be offset from the hub position) is given
by:
r2

V = V0 1 "e

2W2

47 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

where V0 is the undisturbed wind speed, " is the fractional centre line velocity deficit, and W
is the width of the wake (the distance from the wake centre line at which the deficit is
reduced to exp(-0.5) times the centre line value).
Two options are provided for defining the velocity deficit " and the wake width W . They
can be defined directly, or they can be calculated by Bladed by specifying the characteristics
of the upwind turbine. In the latter case, an eddy viscosity model of the wake is used,
developed by Ainslie [7.8,7.9] and described in the next section.
7.3.1 Eddy viscosity model of the upwind turbine wake
The eddy viscosity wake model is a calculation of the velocity deficit field using a finitedifference solution of the thin shear layer equation of the Navier Stokes equations in axissymmetric co-ordinates. The eddy viscosity model automatically observes the conservation
of mass and momentum in the wake. An eddy viscosity, averaged across each downstream
wake section, is used to relate the shear stress term in the thin shear equation to gradients of
velocity deficit. The mean field can be obtained by a linear superposition of the wake deficit
field and the incident wind flow. An illustration of the wake profile used in the eddy
viscosity model is shown in Figure 7.1.

Figure 7.1: Wake profile used in the eddy viscosity model


The Navier Stokes equations with Reynolds stresses and the viscous terms dropped gives
[7.10]:

&U
&U
1 &( ruv)
+V
=
&x
&r
r
&r

The turbulent viscosity concept is used to describe the shear stresses with an eddy viscosity
defined by [7.11]:

48 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

(x) = L m (x).U m (x)


and

uv =

&U
&r

Lm and Um are suitable length and velocity scales of the turbulence as a function of the
downstream distance x but independent of r. The length scale is taken as proportional to the
wake width Bw and the velocity scale is proportional to the difference UI Uc across the
shear layer.
Thus the shear stress uv is expressed in terms of the eddy viscosity.
differential equation to be solved becomes:

The governing

&U
&U
&( r&U / &r )
+V
=
&x
&r
r
&r

Because of the effect of ambient turbulence, the eddy viscosity in the wake can not be wholly
described by the shear contribution alone. Hence an ambient turbulence term is included, and
the overall eddy viscosity is given by [7.12]:

= FK 1 B w ( U i

Uc ) +

amb

where the filter function F is a factor applied for near wake conditions. This filter can be
introduced to allow for the build up of turbulence on wake mixing. The dimensionless
constant K1 is a constant value over the whole flow field and a value of 0.015 is used.
The ambient eddy viscosity term is calculated by the following equation proposed by
Ainslie [7.12]:
2

amb

= F. K k . I amb / 100

Kk is the von Karman constant with a value of 0.4. Due to comparisons between the model
and measurements reported by Taylor in [7.13] the filter function F is fixed at unity.
The centre line velocity deficit Dmi can be calculated at the start of the wake model (two
diameters downstream) using the following empirical equation proposed by Ainslie [7.12]:
D mi = 1

Uc
= Ct
Ui

0.05

[(16C t

0.5)I amb /1000]

Assuming a Gaussian wind speed profile and momentum conservation an expression for the
relationship between the deficit Dm and the width parameter Bw is obtained as

49 of 82

Garrad Hassan and Partners Ltd

Bw =

Document: 282/BR/009

ISSUE:011

FINAL

3.56C t
8D m (1 0.5D m )

Using the above equations, the average eddy viscosity at a distance 2D downstream of the
turbine can be calculated. The equations can then be solved for the centre-line deficit and
width parameter further downstream.
Assuming to the Gaussian profile, the velocity deficit a distance r from the wake centreline is
given by:
r
3.56
Bw

D m ,r = exp

Therefore the wake width W used by Bladed is given by:


W = Bw

0 .5
3.56

7.3.2 Turbulence in the wake


If the eddy viscosity wake model is used, it is also possible to calculate the additional
turbulence caused by the wake. The added turbulence is calculated using an empirical
characterisation developed by Quarton and Ainslie [7.14]. This characterisation enables the
added turbulence in the wake to be defined as a function of ambient turbulence Iamb, the
turbine thrust coefficient Ct, the distance x downstream from the rotor plane and the length of
the near wake, xn. The characterisation was subsequently amended slightly by Hassan [7.15]
to improve the prediction, resulting in the following expression:

I add = 5.7C t 0.7 I amb 0.68 ( x / x n )

0.96

in which all turbulence intensities are expressed as percentages. Using the value of added
turbulence and the incident ambient turbulence the turbulence intensity Itot at any turbine
position in the wake can be calculated as

I tot = I amb + I add

The near wake length xn is calculated according to Vermeulen et al [7.16,7.17]:


in terms of the rotor radius R and the thrust coefficient Ct as

xn =

n r0
dr
dx

50 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

where

r0 = R

m=

n=

m +1
2

1 Ct

(1

0.214 + 0.144m 1
0.214 + 0.144m

0.134 + 0.124m

0.134 + 0.124m

and dr/dx is the wake growth rate:

dr
=
dx
dr
dx
dr
dx
and

dr
dx

dr
dx

+
$

dr
dx

+
m

dr
dx

= 2.5I 0 + 0.005 is the growth rate contribution due to ambient turbulence,


$

=
m

(1

m ) 1.49 + m
is the contribution due to shear-generated turbulence,
(1 + m ) 9.76

= 0.012 B

the number of blades and

is the contribution due to mechanical turbulence, where B is


is the tip speed ratio.

7.4 Time varying wind


Various forms of temporal variation of wind speed and direction may be superimposed on the
spatial variations described in Sections 7.1 to 7.3 above.
7.4.1 Single point time history
A look-up table can be used to supply the wind speed and direction as a function of time, at a
defined reference height. Linear interpolation between time points is used. For any
particular point in space, the wind speed is then multiplied by the appropriate correction
factors for wind shear, tower shadow and upwind turbine wake as defined above.
7.4.2 3D turbulent wind
A 3-dimensional turbulent wind field is generated, with statistical properties representative of
real atmospheric turbulence. Section 7.5 describes how the turbulence is generated. It
consists of dimensionless wind speed deviations, defined as + = (V-Vo)/IV0 where V0 is the
mean wind speed and I the turbulence intensity, at a number of grid points on a rectangular
array large enough to encompass the rotor swept area in the vertical and lateral (cross-wind)

51 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

directions, and long enough in the longitudinal (along-wind) direction to allow a simulation
of the desired length as the whole wind field moves past the rotor at the mean wind speed. At
any point in time, the position in the longitudinal direction is calculated. The position in the
lateral and vertical directions is calculated depending on the radial (r) and azimuthal position
( ) of any particular point on the rotor at that time, and 3-dimensional linear interpolation is
then used to calculate the appropriate wind speed deviation +. The actual wind speed is then
given by
V(r, ,t) = V0Fs0 (Fs + I.+(r, ,t)) .FT .FW
where
Fs0 is the wind shear factor from the reference height (for mean speed V0 ) to the hub height,
Fs is the wind shear factor from the hub height to the point (r, ),
FT is the tower shadow factor for the point (r, ), and
FW is the upwind turbine wake factor for the point (r, ).
7.4.3 IEC transients
The transient variations of wind speed, shear and wind direction defined in the international
standard for the safety of wind turbine systems, IEC 1400-1 [7.1, 7.7], may be simulated with
Bladed. Transient changes in each of the following quantities may be independently
simulated, each with its own parameter values:

Wind speed
Wind direction
Horizontal shear (linear variation of wind speed from one side of the rotor to the other)
Vertical shear (linear variation of wind speed from bottom to top of the rotor)

Each may be either a half-wave transient or a full-wave transient. The transients are
sinusoidal, with a more complex shape defined in edition 2 of the standard [7.7]. The
parameters needed to define each transient are the starting value Y0, the start time t0, the
duration T, and the amplitude A. These parameters are illustrated in Figure 7.2.
12.5

Half wave

Y0 + 12
A

11.5

11

Full wave
10.5

Y0

10

IEC edition 2

9.5

9
-0.2

t00

0.2

0.4

0.6

Time

0.8

t0 1+ T

1.2

Figure 7.2: Definition of IEC sinusoidal transients

52 of 82

Garrad Hassan and Partners Ltd

The actual wind speed at radius r, azimuth

Document: 282/BR/009

ISSUE:011

FINAL

and time t is then given by:

V(r, ,t) = (V0Fs0 Fs + Vtrans) .FT .FW


where V0 is the starting wind speed at the reference height, Vtrans is the combined effect of the
wind speed and horizontal and vertical shear transients, and other parameters as defined in
Section 7.4.2.

7.5 Three dimensional turbulence model


The wind simulation method adopted in Bladed is based on that described by Veers [7.3]. The
rotor plane is covered by a rectangular grid of points, and a separate time history of wind
speed is generated for each of these points in such a way that each time history has the correct
single-point wind turbulence spectral characteristics, and each pair of time histories has the
correct cross-spectral or coherence characteristics.
Calculations using such a turbulent wind field will take into account the crucially important
'eddy slicing' transfer of rotor load from low frequencies to those associated with the
rotational speed and its harmonics. This 'eddy slicing', associated with the rotating blades
slicing through the turbulent structure of the wind, is a significant source of fatigue loading.
The wind speed time histories may, in principle, be generated from any user-specified autospectral density and spatial cross-correlation characteristics. A choice of two different models
of atmospheric turbulence has been provided. These are the von Karman and the Kaimal
models. Both models are generally accepted as good representations of real atmospheric
turbulence, although they use slightly different forms for the autospectral and cross-spectral
density functions. The von Karman model can be used either to generate just the longitudinal
component of turbulence, or to generate all three components if required. Two versions of the
von Karman model are available: the basic model, given in [7.4] and described in Section
7.5.1, and the improved model, described in Section 7.5.2, which is based on more up-to-date
information [7.5, 7.6].
The Kaimal model in Bladed gives only the longitudinal component of turbulence.
It should be remembered, of course, that all these models tend to be based largely on
observations for flat land sites.
7.5.1 The basic von Karman model
The autospectral density for the longitudinal component of turbulence, according to the von
Karman model, is given in [7.4] as

nSuu (n)

, 2u

4n~u
(1 + 70.8n~u2 )5/ 6

where Suu is the auto-spectrum of wind speed variation, n is the frequency of variation, , u is

the standard deviation of wind speed variation and ~


n u is a non-dimensional frequency
parameter given by:

53 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

n xLu
n~u =
U
Here xLu is the length scale of longitudinal turbulence and U is the mean wind speed.
If the three-component model is selected, the corresponding spectra for the lateral (v) and
vertical (w) components are:

nSii (n)

, i2

4n~i (1 + 755.2n~i2 )
=
(1 + 282.3n~i2 ) 11/ 6

where

n x Li
n~i =
U
and i is either v or w.
Associated with the von Karman spectral equations is an analytical expression for the
cross-correlation of wind speed fluctuations at locations separated in both space and time,
derived assuming Taylor's frozen turbulence hypothesis. Accordingly for the longitudinal
component at points separated by a distance "r perpendicular to the wind direction, the
coherence Cu ("r,n), defined as the magnitude of the cross-spectrum divided by the autospectrum, is:

Cu ( "r , n) = 0.994( A5/ 6 (- u )

1
2

- u 5/ 3 A1/ 6 (- u ))

Here Aj(x) = xj Kj(x) where K is a fractional order modified Bessel function, and

nLu ( "r , n)
"r
- u = 0.747
1 + 70.8
Lu ( "r , n)
U

The local length scale Lu("r,n) is defined by:


L u ("r, n ) = 2MIN(1.0,0.04n

2/3

( y L u "y) 2 +( z L u "z) 2
"y 2 + "z 2

where "y and "z are the lateral and vertical components of the separation "r, and yLu and zLu
are the lateral and vertical length scales for the longitudinal component of turbulence.
For the lateral and vertical components, the corresponding equations are:

Ci ( "r , n) =

0.597
4.781# i2 A5/ 6 (- i )
2
2.869# i 1

54 of 82

A11/ 6 (- i )

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

where

nLi ( "r , n)
"r
- i = 0.747
1 + 70.8
Li ( "r , n)
U
and

#i =

- i Li ( "r , n)
"

for i = v or w.
In this case the local length scales are given by:
L v ("r, n ) = 2MIN(1.0,0.05n

2/3

( y L v "y / 2) 2 +( z L v "z) 2

"y 2 + "z 2

and
L w ("r, n ) = 2MIN(1.0,0.2n

1/ 2

( y L w "y) 2 +( z L w "z / 2) 2
"y 2 + "z 2

The three turbulence components are assumed to be independent of one another. This is a
reasonable assumption, although in practice Reynolds stresses may result in a small
correlation between the longitudinal and vertical components near to the ground.
7.5.2 The improved von Karman model
The improved von Karman model [7.5] attempts to rectify some deficiencies of the basic
model at heights below about 150m. The autospectral density for the longitudinal component
of turbulence is given by:

nS uu (n)

, u2

= .1

2.987n~u / a
2
1 + ( 2 n~u / a )

5/ 6

+ .2

1294
. n~u / a

(1 + (

2
n~u / a )

5/ 6

F1

where Suu is the auto-spectrum of wind speed variation, n is the frequency of variation, ,u is
the standard deviation of wind speed variation and n~u is a non-dimensional frequency
parameter given by:

n xLu
n~u =
U
Here xLu is the length scale of longitudinal turbulence and U is the mean wind speed.
If the three-component model is selected, the corresponding spectra for the lateral (v) and
vertical (w) components are:

55 of 82

Garrad Hassan and Partners Ltd

nSii (n)

, i2

= .1

Document: 282/BR/009

2.987(1 + (8 / 3)(4 n~i / a ) 2 )(n~i / a )

2
1 + ( 4 n~i / a )

11/ 6

+ .2

ISSUE:011

1294
. n~i / a
2
1 + (2 n~i / a )

FINAL

5/ 6

F2

where

n x Li
n~i =
U
and i is either v or w.
The five additional parameters a, . 1, . 2, F1 and F2 are defined as follows:

[
]
F = 1 + 2.88 exp[ 0.218( n~ / a ) ]
F1 = 1 + 0.455 exp 0.76(n~u / a )

0.8

0. 9

. 2 = 1 .1
.1 = 2.357a 0.761

a = 0535
.
+ 2.76(0138
.
A) 0.68
where

A = 0115
. [1 + 0.315(1 z / h) 6 ]2 / 3
Here z is the height above ground, and h is the boundary layer height obtained from:

h = u * / (6 f )

f = 2 sin(

(the Coriolis parameter:


the earth, and

u = ( 0.4U 34.5 f . z ) / ln( z / z0 )


z0 = surface roughness length

is the angular speed of rotation of

is the latitude)

The turbulence intensities of the three components of turbulence are also defined for the same
choice of z, z0, U and , as follows:

- = 1 6 f . z / u*
p = -16
,u =

7.5-( 0538
.
+ 0.09 ln( z / z0 )) p u *
1 + 0156
. ln u * / f . z0

Iu = , u / U

I v = I u 1 0.22 cos4

(the longitudinal turbulence intensity)

z
2h

(the lateral turbulence intensity)

56 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

I w = I u 1 0.45 cos4

z
2h

ISSUE:011

FINAL

(the vertical turbulence intensity)

although these may be changed by the user for any particular simulation.
The nine turbulence length scales are also defined, as follows:

Lu =

A1.5 , u / u * z
2.5Kz1.5 (1 z / h) 2 (1 + 5.75z / h)

(
= 05
. L (1

(
0.68 exp(

Lu = 0.5x Lu 1 0.46 exp 35( z / h) 1.7

Lu

))
))

35( z / h)1.7

Lv = 05
. x Lu (, v / , u ) 3
x
Lw = 0.5x Lu ( , w / , u ) 3
y
Lv =2 y Lu (, v / , u ) 3
z
Lv = zLu ( , v / , u ) 3
y
Lw = yLu (, w / , u ) 3
z
Lw =2 z Lu (, w / , u ) 3
x

where

K z = 019
.
( 019
.
K0 ) exp B( z / h)

K0 = 0.39 / R 0.11
B = 24 R 0.155
N = 124
. R 0.008
u*
R=
f . z0
Associated with the von Karman spectral equations is an analytical expression for the
cross-correlation of wind speed fluctuations at locations separated in both space and time,
derived assuming Taylor's frozen turbulence hypothesis [7.6]. Accordingly for the
longitudinal component at points separated by a distance "r perpendicular to the wind
direction, the coherence Cu ("r,n), defined as the magnitude of the cross-spectrum divided by
the auto-spectrum, is:

Cu ( "r , n) = 0.994( A5/6 (- u )

1
2

- u 5/3 A1/ 6 (- u ))

Here Aj(x) = xj Kj(x) where K is a fractional order modified Bessel function, and

-i =

0.747 "r
2 Li

2 n"r
+ c
U

for i = u

57 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

The local length scale Lu("r,n) is defined by:


( y L u "y) 2 + ( z L u "z) 2

L u ("r, n ) =

"y 2 + "z 2

while

c = max(10
. ,
with

16
. ( "r / 2 Lu ) 0.13

- 0b

b = 0.35( "r / 2 Lu )

0.2

and

-0 =

0.747 "r
2 Lu

2 n"r
+
U

"y and "z are the lateral and vertical components of the separation "r, and yLu and zLu are the
lateral and vertical length scales for the longitudinal component of turbulence.
For the lateral and vertical components, the corresponding equations are:

Ci ( "r , n) =

0.597
4.781# i2 A5/ 6 (- i )
2
2.869# i 1

A11/ 6 (- i )

for i = v,w

where -i is defined as above for i = v, w, and

#i =

- i 2 Li ( "r , n)
"

In this case the local length scales are given by:


L v ("r, n ) =

( y L v "y / 2) 2 +( z L v "z) 2
"y 2 + "z 2

and
L w ( "r, n ) =

( y L w "y) 2 +( z L w "z / 2) 2
"y 2 + "z 2

The three turbulence components are assumed to be independent of one another. This is a
reasonable assumption, although in practice Reynolds stresses may result in a small
correlation between the longitudinal and vertical components near to the ground.

58 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

7.5.3 The Kaimal model


The autospectral density for the longitudinal component of turbulence, according to the
Kaimal model, is:

nSuu (n)

, 2u

4n~u
(1 + 6.0n~u )5/ 3

where Suu is the auto-spectrum of wind speed variation, n is the frequency of variation, , u is
the standard deviation of wind speed variation and n~u is a non-dimensional frequency
parameter given by:

n L1
n~u =
U
Here L1 = 2.329 xLu where xLu is the length scale of longitudinal turbulence, and U is the
mean wind speed as before.
A simpler coherence model is used in conjunction with the Kaimal model. With the same
notation as in Section 7.5.1, the coherence is given by

C ( "r , n) = exp

8.8"r

n
U

012
.
+
L( "r , n)

7.5.4 Compatibility with IEC 1400-1


The turbulence model defined in the IEC standard 1400-1 [7.1] assumes isotropic turbulence.
In this case, we have xLu = 2 yLu = 2 zLu and n-2/3 high frequency modification to the local
length scale L("r,n) is not applicable. The above relationships are in fact equivalent to the
IEC 1400-1 definition for frequencies below 0.008 Hz provided xLu = 2 yLu = 2 zLu.
7.5.5 Using 3d turbulent wind fields in simulations
The following points should be noted when using these turbulent wind fields for wind turbine
simulations:
The length of the wind field, Lwind, must be sufficient for the simulation to be carried out.
For a simulation of T seconds at a mean wind speed of U m/s, Lwind must be at least UT +
D metres where D is the turbine diameter (the extra diameter is needed in case the turbine
is yawed with respect to the mean wind direction).
The width and height of the wind field must evidently be sufficient to envelope the whole
rotor, i.e. at least equal to the rotor diameter.
A grid of about 7x7 points to cover the rotor plane is generally sufficient. Clearly the
number of points required to achieve suitable resolution of the spatial turbulent variations
will depend on the ratio of the turbulence length scales used to the rotor diameter.
If a simulation uses only a part of a turbulent time history, the mean wind speed and
turbulence intensity for that part of the time history may not be the same as for the whole
time history, and therefore may not match the mean wind speed and turbulence intensity
59 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

which was specified for the simulation since this assumes that the whole time history will
be used. Note also that a turbulence history of length Lwind with an along-wind step size of
"L would be calculated at Lwind/"L points in the along-wind direction. This must be a
power of two for efficient calculation, since Fast Fourier Transform techniques can then be
used. If it is not a power of two, then the spacing "L will automatically be decreased to
make Lwind/"L a power of two.
Different time histories with the same turbulence characteristics can be generated by
changing the random number seed.
A sinusoidal half- or full-wave wind direction transient as described in Section 7.4.3 may
be superimposed on the turbulent wind field. This is intended for use with turbulent wind
fields when only the longitudinal component has been generated, to ensure that some yaw
error occurs during the simulation. Using all three components of turbulence should give a
more realistic variation of yaw error.

60 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

8. MODELLING WAVES AND CURRENTS


For wind turbines sited offshore, the fatigue loads and extreme loads experienced by the
tower are strongly dependent on the action of waves and currents on the tower base. For
fatigue load calculations in particular it is important to couple the wind and wave load
calculations so that both aerodynamic and hydrodynamic damping act together to moderate
tower movement.
For fatigue load calculations, Bladed creates a series of irregular waves based on linear Airy
theory. The amplitude and frequency content of these waves are specified by the user in
terms of a power spectral density function. This may be either:
the standard JONSWAP / Pierson-Moskowitz function, or
a user-defined function.
For extreme load calculations, a regular wave train may be defined. The kinematics of this
wave are calculated using stream function theory.

8.1 Tower and Foundation Model


Offshore wind turbines are most likely to be installed in relatively sheltered inshore
conditions, where the sea depth is in the range 5m to 25m. Bladed assumes that the tower is
fixed to the sea bed as a simple monopile as shown in Figure 8.1 below. The tower may be
defined over the full depth (Figure 8.1a) or above a rigid base (Figure 8.1b). In both cases,
the turbine structure is regarded as being transparent to the waves, implying that both tower
and base are slender in comparison to the wavelength.

a) Simple Monopile

b) Monopile with narrow base

Figure 8.1: Assumed base structures


As for onshore cases, the tower is assumed to have a circular cross-section and may be
tapered. Foundation translational and rotational stiffnesses may also be specified.

61 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

8.2 Wave Spectra


To create an irregular wave train for fatigue load calculations the user must specify a suitable
wave spectral formula S! ( f ) . This function will depend on the location of the turbine being
modelled and the prevailing meteorological and oceanographic conditions. Bladed allows the
wave spectrum to be specified in one of two ways: as a JONSWAP / Pierson-Moskowitz
spectrum or as a user-defined look-up table.
8.2.1 JONSWAP / Pierson-Moskowitz Spectrum
There are several different versions of the JONSWAP formula. The version used is based on
an expression by Goda [8.1].
S! ( f ) = $

2
2 H s Tp

f
fp

f
125
.
fp

exp

#.

where f is the wave frequency (in Hz), H s is the significant wave height, Tp is the peak
spectral period, f p = 1 Tp , # is the JONSWAP peakedness parameter,

$2 =

0.0624

. = exp 0.5

and

0.185
1.9 + #

0.230 + 0.0336#

, = 0.07 for f

f
fp

fp

, = 0.09 for f > f p


The Pierson-Moskowitz spectral density function may be regarded as a special case of the
JONSWAP spectrum with # = 1.0 :
S! ( f ) =

0.3123H s2 Tp

f
fp

exp

f
1.25
fp

If the JONSWAP / Pierson-Moskowitz option is selected, the user is required to enter values
for H s , Tp and # .
8.2.2 User-defined Spectrum
A user-defined spectrum may be entered in the form of a look-up table. Up to 100 pairs of
S! ( f ) and f may be entered. The values of S! ( f ) at the lowest and highest frequencies
entered should be zero. At frequencies between the specified values of f , values of S! ( f )
are linearly interpolated.
62 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

8.3 Upper Frequency Limit


Waves which have wavelengths much smaller than the diameter of the tower do not
contribute to the net force because regions of positive and negative velocity are experienced
by the tower at the same time. Applied forces are calculated from the wave particle
kinematics at the tower centreline only and so the calculations specifically exclude these high
frequency components. The frequency cut-off is based on experimental work by Hogben and
Standing [8.2] which show that the applied force on a cylinder falls off rapidly when the
wave
number
exceeds
1 / radius. Therefore:
S! ( f ) = 0 for k >

1
radius

The radius is taken as the minimum tower radius between the sea bed and a height of 3
standard deviations of the wave elevation above the mean water level. At any instant, the
wave elevation has a probability of 99.85% of being within this range.

8.4 Wave Particle Kinematics


For both the fatigue and extreme wave load calculations, wave particle kinematics are based
on linear Airy theory. The following equations describe the wave particle velocity vector
u w = u wx , u wy , u wz , the corresponding acceleration vector u& w = u& wx , u& wy , u& wz , the
hydrodynamic component of the pressure p and the water surface elevation ! for a regular
wave of height H and period T at the point ( x , y , z ) :

u wx =

H
cos w cosh k (d + z ) cos($
2 sinh( kd )

t)

u wy =

H
sin w cosh k (d + z ) cos($
2 sinh( kd )

t)

u wz =

H
sinh k (d + z ) sin($
2 sinh( kd )

u& wx =
u& wy =
u& wz =

t)

H
cos w cosh k (d + z ) sin($
2 sinh( kd )

H
sin w cosh k (d + z ) sin($
2 sinh( kd )

H
sinh k (d + z ) cos($
2 sinh( kd )

p=

gH
cosh k (d + z ) cos($
2 cosh( kd )

!=

H
cos($
2

t)
t)

t)
t)

t)

where
= 2 f is the angular wave frequency, f is the wave frequency, t is time, d is the
water depth (assumed to be constant), is the water density, g is the acceleration due to
gravity and
63 of 82

Garrad Hassan and Partners Ltd

$ = kx cos w

Document: 282/BR/009

ISSUE:011

FINAL

ky sin w

where w is the direction from which waves arrive at the tower. The wave number k is found
as the solution to the dispersion relation:
2

= gk tanh kd

The co-ordinate system used for the wave and current calculations is a right-handed Cartesian
system in which the xy plane is horizontal with the x-axis pointing to the North, the y-axis
pointing to the West and the z-axis pointing vertically upwards. The origin of the co-ordinate
system lies where the tower centre line intersects the mean water level. Angles are defined
relative to the x-axis (North) and increase positively toward the East.
For the calculation of regular extreme waves, the above equations are used directly to
calculate the wave particle kinematics at each submerged tower station. For fatigue load
calculations, however, it is necessary to calculate an irregular (i.e. random, non-repeating)
series of waves. This is achieved using the filtered white noise shift register procedure
described in section 8.6 below.

8.5 Wheeler Stretching


A limitation of Airy theory is that it only defines wave particle kinematics up to the mean
water level (z = 0). The theory can be extended above the mean water surface, up to the level
of the wave crest, by using the Airy formulae with positive values of z. However this
approach causes calculation difficulties and is known to over-estimate particle velocities and
accelerations in the crest region and to underestimate velocities and accelerations in the
troughs. To avoid these difficulties, Bladed uses Wheeler stretching [8.3] to take account of
the forces acting between mean water level and the instantaneous free surface. Experimental
results by Gudmestad [8.4] indicate that Wheeler stretching provides satisfactory estimates of
particle kinematics in the free surface zone in deep water.
Wheeler stretching assumes that particle motions calculated using Airy theory at the mean
water level should actually be applied at the instantaneous free surface. Airy particle motions
calculated at locations between the sea-bed and mean free surface are shifted vertically to
new locations in proportion to their height above the sea bed. Airy wave particle kinematics
calculated at a vertical location z are therefore applied to a new location z defined by:
z =

d + ! (t )
d

+ ! (t )

where ! (t ) is the surface elevation above the location in question.

8.6 Simulation of Irregular Waves


During an offshore simulation in which waves are specified, the following records are
synthesised:

64 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

Wave elevation at the tower centre-line,


Wave particle velocities, accelerations and dynamic pressures at various points on the
structure,
Wave forces on the submerged tower.
For irregular waves, these records are created by the digital filtering of pseudo-random white
noise. A single white noise record is used, together with a different filter for each time
history to be generated. Because each filter introduces the correct amplitude variation and
phase shift, the resulting output time histories display the correct amplitude and phase
relationships to each other. Unlike the generation of turbulent wind records, which are
generated and written to a file before running the simulation, wave data are generated as the
simulation proceeds.
The relationship between the parameter of interest (i.e. the wave particle velocity at the first
tower station, the particle acceleration at the sea bed etc.) and the water surface elevation is
defined in terms of a complex function of the wave frequency known as a Response
Amplitude Operator (RAO). It is represented as a complex number of the form:
RAOr = Rr e i0 r

The filters used to process the pseudo-random white noise are Finite Impulse Response (FIR)
filters and are defined in terms of their frequency transforms. The transformed filter for
response r is given by:
zm,r = Rr ( f m )

m ,r

S ( f m )"f
4N

exp i0 r ( f m )

= z m ,r

where
f m = m"f
"f = f max N

and m is in the range 0

N.

The filter weights are then obtained as the transform of the expression:
wn ,r =

m = N +1

z m,r exp

imn
N

Having generated the filter functions for each parameter at each required location, time
histories are generated using a shift-register technique. Firstly an N-element array of
normally-distributed random numbers is created. The random numbers are generated by
converting the output of a simple random number generator to a normally distributed deviate
with zero mean and unit variance using the Box-Muller method. For each filter function in
turn, the N filter weights are multiplied by the values of the equivalent elements in the
random number array and the N products are then summed to give the value of the property at
one particular instant in time. To calculate the value of the property at the next time step, the
elements of the random number array are shifted one place higher in the array, a new

65 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

random number is introduced at element 1 and the multiplication and summation process is
repeated.

8.7 Simulation of Regular Waves


If Extreme Deterministic Waves are selected from the Waves panel, the wave kinematics
are calculated using stream function theory. This method is more accurate than linear wave
theory in cases where the wave height is a significant proportion of the mean water depth.
The method may even be used to model waves with amplitudes close to the breaking wave
limit. In cases when currents are specified in addition to regular waves (see section 8.8), the
wave calculation takes proper account of the influence of the current profile on the wave
kinematics. The non-linear regular wave calculations within Bladed are based on original
coding by Chaplin [8.5].
Regardless of whether current components are specified, Bladed first solves the wave
equation using stream function theory for the case of no currents. Stream function theory was
first applied to wave modelling by Dean [8.6 & 8.7] who developed the following form of
stream function:
X1
z+
T

( x, z ) =

where

N
n=2

X n sinh(nk ( z + d )) cos(nkx)

X 1 = wavelength
X n +1 =

and N is the order of the stream function solution.


as defined above satisfies the requirements that (i) the shape of the
The stream function
free surface is compatible with the motion of the water just below it (the Kinematic Free
Surface Boundary Condition), (ii) the flow is periodic, and (iii) the flow is compatible with
the presence of a horizontal sea bed at the specified depth. The values of X n are determined
by a least-squares method to satisfy the additional requirements that (i) the pressure on the
free surface is uniform (the Dynamic Free Surface Boundary Condition), and (ii) the required
wave height is obtained. As implemented within Bladed, the order of the solution, N, is
automatically chosen based on the input values of wave height, period and mean water depth.
Once the stream function solution has been obtained, the horizontal and vertical velocities (in
the absence of a current) are calculated using the relations:

u=

&
&y

and

v=

&
&x

and the dynamic pressure is calculated using Bernoullis equation.


In cases where a current profile is specified the flow is in general rotational and the wave
solution must be modified. The method used follows the approach developed by Dalrymple
[8.8 & 8.9] and is based on coding by Chaplin [8.5]. It is assumed that the relationship

66 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

and stream function


is the same for the combined flow as for the
between the vorticity
undisturbed current when viewed from a reference frame moving at such a speed that total
flow rate is the same as that in the x,y frame. This requirement can be stated mathematically
as:

&2
&2
+
=
&x 2 &y 2

= f( )

In Bladed, the stream function


is computed at discrete points in the x,y plane using a
finite-difference calculation scheme. The most difficult feature of this approach is that the
location of the free surface is not known in advance. A regular grid of points in the x,y plane
would therefore have awkward intersections with the free surface profile, which must itself
be calculated as part of the computation. To overcome this difficulty, Dubreil-Jacotins
method is used to transform the problem from the x,y plane to the x, plane, with y as the
field variable. The position of the free surface is now defined along the upper boundary of a
plane. Treating x and
as the independent parameters, the
rectangular grid in the x,
velocity components are now given by:

u=

1
&y

&y
and

&

v=

&y

&x

&

The accuracy of the solution relies on a sufficiently fine mesh in the x, plane to resolve
the structure of the flow and to allow the evaluation of derivatives on the boundaries of the
computational domain, particularly at the free surface. For this purpose a regular grid in the
x, plane is rather inefficient and therefore a stretched grid is employed which is finer near
the free surface than the sea bed.
After solving the finite difference relations on this plane, the flow velocities are calculated
using the equations above and dynamic pressures are calculated using Bernoullis equation.
Reference [8.5] should be consulted for further details of this method.

8.8 Current Velocities


Bladed allows current velocities to be calculated based on three current profiles, either
separately or in combination:
a near-surface (wind/wave generated) current: u cw
a sub-surface (tidal and thermo-saline) current: u cs
a near-shore (wind induced surf) current: u cn
These three velocity vectors have the form:

67 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

u cw = ucw ( z ) (cos cw , sin cw ,0)


u cs = ucs ( z ) (cos cs , sin cs ,0)

u cn = ucn ( z ) (cos cn , sin cn ,0)

where cw , cs and cn are the directions from which the three current components arrive
at the tower. Components of the calculated current velocities are then combined linearly:
u c = u cw + u cs + u cn

8.8.1 Near-Surface Current


The near-surface current velocity profile is of the form:
u cw ( z ) = 2 ( z ) us ( z10 )

where us ( z10 ) is an input parameter, representing the mean wind speed at a height 10m above
the mean water surface. 2 ( z ) is given by the formulae:
z
if 15m
15
2 ( z ) = 0.0 if z < 15m

2 ( z ) = 0.01 1

0m

8.8.2 Sub-Surface Current


The sub-surface current velocity profile is of the form:
ucs ( z ) =

z+d
d

17

u s 0 ( z = 0)

for 0 4 z 4 d , where d is the water depth and u s0 ( z = 0) is an input parameter equal to the
velocity at the sea surface.
8.8.3 Near-Shore Current
The near-shore current velocity has a uniform profile, independent of depth. The design
velocity at the location of the breaking wave is defined as:
ucn = 2 s gH B

where g is the acceleration due to gravity, s is the beach slope and H B is the breaking wave
height given by:
HB =

b
a
1
+
d B gTB2

68 of 82

Garrad Hassan and Partners Ltd

where:

Document: 282/BR/009

a = 44 1 exp( 19 s)

b = 1.6 1 + exp( 19 s)

ISSUE:011

FINAL

d B is the water depth at the location of the breaking wave and TB is the period of this wave.
For very small beach slopes H B may be estimated using the formula H B = 0.8d B .

8.9 Total Velocities and Accelerations


The wave particle velocity and acceleration vectors at a particular location ( x , y , z ) in the
wave field at time t, obtained from the white noise filtering procedure, are denoted by u w and
u& w . The total current velocity vector at the same location is u c and the velocity and
acceleration of the tower structure itself are u s and u& s .
The total velocity u t and acceleration u& t of the fluid relative to the structure at this location
and time are therefore:
ut = uw + uc
u& t = u& w u& s

us

8.10 Applied Forces


Having evaluated the total particle kinematics relative to the tower, the resulting forces are
calculated as the sum of two components:
Drag and inertia forces calculated using the relative motion form of Morisons equation,
Longitudinal pressure forces.
These forces are then used to calculate the tower modal forces as described in Section 3.2.2.
8.10.1 Relative Motion Form of Morisons Equation
To calculate the forces on the tower, the monopile is approximated by 10 cylindrical subelements of equal height. The forces on each sub-element, acting normal to the cylinder axis,
are calculated using the relative motion form of Morisons equation:
F = ( Cm

1)

D2
u& t +
4

D2
1
Lu& w + Cd DLut ut
4
2

where F is the normal force on a segment of cylinder of length L and diameter D,


water density, Cm is the inertia coefficient and Cd is the drag coefficient.

is the

8.10.2 Longitudinal Pressure Forces on Cylindrical Elements


Morisons equation gives the force on the cylindrical element normal to the elements axis. In
situations where the monopile is strongly tapered, pressures acting longitudinally on the
changing cross-sectional area may cause a significant vertical force to act on the pile.

69 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

The added mass acting in the longitudinal direction is very small and so the longitudinal
forces are estimated using the hydrodynamic pressure in the ambient wave field acting over
the change in cross-sectional area of the tower between the top and bottom faces of each subelement. For a tower with diameter Da at the top of a sub-element and diameter Db at the
bottom, the longitudinal force acting on this portion of the tower is:
F=

(D

2
a

Db2

)p

No pressure force is included where the end of the tubular member passes through the free
surface or terminates at the sea-bed.

70 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

9. POST-PROCESSING
Bladed includes an integrated post-processing facility which allows the results of wind
turbine calculations to be processed further in various ways. The theory behind these postprocessing calculations is described in this section.

9.1 Basic statistics


The following basic statistical properties of a signal are calculated:
Minimum
Maximum
Mean

MIN(x)
MAX(x)
x

Standard deviation

,=

Skewness

(x

x )3 / , 3

Kurtosis

(x

x )4 / , 4

(x

x )2

9.2 Fourier harmonics, and periodic and stochastic components


Wind turbine loads consist of both periodic and random or stochastic components. The
periodic components of loads result from effects which vary as a function of rotor azimuth,
such as gravitational loads, tower shadow, yaw misalignment, wind shear etc. The stochastic
components result from the random nature of wind turbulence. In understanding the loads on
a wind turbine it is often useful to separate out the periodic and stochastic parts of a load time
history, and future to analyse the periodic part in terms of the harmonics of the fundamental
rotational frequency.
The periodic part of a signal is obtained by binning the signal against rotor azimuth. The
number of azimuth bins may be specified by the user, otherwise it is calculated from the first
two azimuth values in the time history. These are used to define the azimuth bin width,
which is then adjusted to an exact sub-multiple of a revolution.
The number of azimuth bins must be compatible with the sampling interval of the time
history. If too many bins are used, it is possible for some of them to be empty, in which case
the calculation will not proceed.
Having obtained the periodic component of the signal, the Fourier harmonics are obtained by
means of a discrete Fourier transform, after first increasing the number of bins by two to four
times using linear interpolation.
The stochastic component of the signal is obtained for each time point by subtracting the
periodic component calculated from the azimuth at that time point. Linear interpolation is
used between azimuth bins.

71 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

9.3 Extreme prediction


The prediction of the extreme loads which are likely to be encountered by a wind turbine
during its lifetime is clearly a crucially important part of the design process. It is common
practice to base the prediction of these extreme loads on deterministic load cases, in which
the wind turbulence is represented in terms of discrete gusts with amplitudes and rise times as
specified by design standards and certification rules. Discrete gusts can be modelled with
Bladed as described in Section 7.4.3.
An alternative approach, which avoids the problem of the rather arbitrary nature of these
discrete gusts, is based on probabilistic techniques, with the stochastic nature of the loads due
to wind turbulence represented by means of a probability distribution. Although this
approach has been used for many years for the evaluation of extreme loads on buildings and
similar structures, its application to wind turbine loads is relatively rare. The analysis
involved in applying it to an operational wind turbine is rather more complicated since the
probability distribution of the combined stochastic and deterministic load components must
be considered.
Any particular wind turbine loading can be expressed as
y(t) = z(t) + x(t)
where z and x represent the periodic and stochastic parts of the load respectively (see Section
9.2). It is generally a good approximation to assume that the stochastic part of the load is
Gaussian, so its probability distribution is:

p( x ) =

,x 2

x 2 / 2, 2x

where ,x is the standard deviation of x. For such a signal, Rice [9.1] has derived the
probability distribution of signal peaks as:

p$ ( x ) =

1 #2

,x 2

- 2 / 2 (1 # 2 )

-#
e
2, x

-2 / 2

1 + erf

#2

where

- = x /, x
# = 50 /5m

50 =

M2
M0

(the zero up-crossing or apparent frequency)

5m =

M4
M2

(the frequency of peaks)

72 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

Mi =

f i H ( f )df

(the ith spectral moment)

f = frequency (Hz)
H ( f ) = power spectral density (see next section for calculation details), and
erf () = error function.
Knowing the probability distribution of peaks for such a process, the probability distribution
of extremes can then be deduced. For the extreme of the signal in a given period to be x, one
peak must have this value and all other peaks in the period must have a lesser value. The
probability distribution can be written
N
p$$ (-) = Np$ (-)(1 Q(-))

where
6

Q(-) = p$ (-) d- , and


-

N = Number of peaks in the period.


Davenport [9.2] combined this with Rices equation to give the following analytical
expression for the probability distribution of extremes:

p$$ (-) = - e
where
2

= 5 0 T e - / 2 and
T = time period.
The mean of this distribution is

-ext = . +

$
.

where

. = 2 ln(5 0 T ) and
$ = 0.5772 (Eulers constant).
As the term 5 0 T increases, the distribution of extremes has a larger mean and becomes very
narrow.
For an operational wind turbine whose loads are a combination of stochastic and periodic
components, Madsen et al [9.3] proposed an approach based on Davenports model of the
stochastic signal, with the assumption that the extremes in the total signal occurred at minima
and maxima of the periodic component. This allows the periodic time history to be idealised
as a square waveform as follows:

73 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

z
t1

zmax

t2

zmean

Time
t3

zmin
T0

The resulting expressions for the mean and standard deviation of the extreme distributions
are:
For extreme maxima:

ye max = z max + , x . 1 +

, e max = , x

$
.1

6. 1

where

. 1 = 2 ln( 15 0T )
1 =

t1
=
T0 ( zmax

, 2z
zmean )( z max

zmin )

while for extreme minima:

ye min = zmin
, e min = , x

, x .3 +

$
.3

6. 3

where

. 3 = 2 ln( 35 0T )
3

t
= 3 =
T0 ( zmean

, 2z
zmin )( z max

z min )

Here , z is the standard deviation of the periodic component z. The time period T should be
taken as the total time for which the condition being modelled will be experienced during the
lifetime.

74 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

9.4 Spectral analysis


Bladed allows the calculation of auto-spectral density, cross-spectral magnitude and phase,
transfer functions and coherence functions.
All calculations involving spectral analysis use a Fast Fourier Transform technique with
ensemble averaging. To perform the spectral analysis, the signal is divided into a number of
segments of equal length, each of which contains a number of points which must be a power
of 2. The segments need not be distinct, but may overlap. Each segment is then shaped by
multiplying by a window function which tapers the segment towards zero at each end. This
improves the spectrum particularly at high frequencies. A choice of windowing functions is
available. Optionally, each segment may have a linear trend removed before windowing,
which can improve the spectral estimation at low frequencies. The final spectrum is obtained
by averaging together the resulting spectra from each segment, and scaled to readjust the
variance to account for the effect of the window function.
The information required is therefore as follows:
Number of points: the number of datapoints per segment. This must be a power of 2: if it is
not, it is adjusted by the program. The maximum allowed is 4096. The larger the number of
points, the better will be the frequency resolution, which may be important especially at low
frequencies. However, choosing fewer points may result in a smoother spectrum because
there will be more segments to average together. If in doubt, 512 is a good starting point.
Percentage overlap: the overlap between the segments. This must be less than 100%. 50%
is often satisfactory, although 0% may be more appropriate if a rectangular window is used.
Window: a choice of five windowing functions is provided:
(a) rectangular (equivalent to not using a window)
(b) triangular: 1 2 f 1
(c) Hanning:
(d) Hamming:

(1 cos(2 f )) / 2
0.54 0.46 cos(2 f )

(e) Welch:

1 (2 f

1) 2

where f is the fractional position along the segment (0 at the start, 1 at the end). One of the
last three windows (which are all quite similar) is recommended.
Trend removal is usually desirable.

9.5 Probability, peak and level crossing analysis


These calculations work by binning values. The range and size of the bins to be used are
calculated by the program, unless they have been supplied by the user.
The probability density analysis simply bins the signal values. From the probability density
function it also calculates the cumulative probability distribution. Also a Gaussian
distribution is calculated for comparison, which has the same mean and standard deviation as

75 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

the signal. There is an option to remove the mean of the signal: this merely moves the mean
of the calculated distribution to zero.
The peak analysis bins only those signal values which are turning points of the signal. Peaks
and troughs are binned separately, so that the probability distribution of each can be output.
For the level crossing analysis, the number of up-crossings and down-crossings are counted at
each of the bin mid-points. The number of crossings per unit time in each direction is output
for each bin mid-point.

9.6 Rainflow cycle counting and fatigue analysis


Bladed offers the possibility of rainflow cycle counting of a stress time history and of
subsequent fatigue analysis based on the cycle count data. A suitable stress time history can
be generated from one or more load time histories by use of the channel combination and
factoring facility provided by the code.
9.6.1 Rainflow cycle counting
Rainflow cycle counting is the most generally accepted method used as the basis of fatigue
analysis of structures. The key advantage of the rainflow cycle counting method is that it is
able to take proper account of stress or strain reversals in the context of a stress-strain
hysteresis loop.
The cycle counting procedure involves the following steps:
The stress history is searched to determine the successive peaks and troughs by
identification of turning points.
The successive peaks and troughs are re-ordered so that the sequence begins with the
highest peak value of the stress history.
The sequence of peaks and troughs is now scanned to determine the rainflow cycles. A
rainflow cycle is only recorded when the range exceeds a user specified minimum range.
The purpose of this user-specified minimum range is to filter out very small cycles where
this is desired.
The mean and range of each rainflow cycle is recorded.
The count of rainflow cycles is binned according to the cycle mean and range values. The
distribution of bins is defined by the user who is required to specify minimum and
maximum values of stress and the number of bins to be used.
The output from the rainflow cycle counting analysis consists of the two-dimensional
distribution of the number of cycles binned on the means and ranges of the cycles.
This calculation can also be extended to generate damage equivalent loads. The user
specifies one or more inverse S-N slopes m (see next section) and a frequency f (typically 1P
for fixed speed machines), and an equivalent load is calculated as the amplitude of a

76 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

sinusoidal load of constant frequency f which would produce the same fatigue damage as the
original signal. The equivalent load is therefore given by:

1 ni Si m

1
m

Tf
where ni is the number of cycles in stress range Si and T is the duration of the original time
history.
9.6.2 Fatigue analysis
As is described above, a complex stress history can be represented in terms of constituent
cycles by use of the rainflow cycle counting technique. The distribution of rainflow cycles is
defined in terms of the number of cycles binned against stress range and mean value.
The basis of the fatigue analysis provided in Bladed is that fatigue failure is predicted to
occur according to the Palmgren-Miner [9.4] linear cumulative damage law. Failure will
occur when the accumulated fatigue damage number is equal to 1.0 as follows::
ni

1N
i

= 1.0

where ni is the number of rainflow cycles of the ith stress range and Ni is the corresponding
number of cycles to failure. The summation is defined as the accumulated damage.
For rainflow cycles of stress range Si, the number of cycles to failure Ni is given by the S-N
curve for the material. The user of Bladed must supply the S-N curve in one of two ways.
The first possibility is that the S-N curve is provided as a log-log relationship of the form:
log S =

1
log k
m

1
log N
m

so that:
N = kS

The user must specify the value of m, the inverse slope of the log S against log N relationship.
The user must also specify the intercept of the log-log relationship, c. The parameter k above
is related to the intercept c by:
k = cm

The second option is for the user to specify the S-N curve as an arbitrary function through the
use of a look-up table.
For a material where the mean stress has an influence on the fatigue damage accumulated,
Bladed offers the option of converting each cycle range to the equivalent range assuming a
zero mean stress value. (A cycle with a zero mean value has a R-ratio of -1, where R is the
77 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

ratio of minimum to maximum stress.) This conversion is performed by means of a Goodman


diagram and the user is required to provide the ultimate tensile strength (UTS) of the
material. Following the conversion, the fatigue analysis proceeds using the Palmgren-Miner
law and user specified S-N curve as described above.
The output from the fatigue analysis consists of the accumulated damage due to the stress
history as well as the two-dimensional distribution of the proportion of the accumulated
damage binned on the means and ranges of the rainflow cycles.

9.7 Annual energy yield


The annual energy yield is calculated by integrating the power curve for the turbine together
with a Weibull distribution of hourly mean wind speeds. The power curve is defined at a
number of discrete wind speeds, and a linear variation between these points is assumed.
The Weibull distribution is defined by:

F (V ) = 1 e

V
cV

where F is the cumulative distribution of wind speed V. Thus the probability density f(V) is
given by

Vk 1
f (V ) = k
e
(cV ) k

V
cV

Here k is the Weibull shape factor, and c is the scale factor. For a true Weibull distribution,
these two parameters are related by the gamma function:

c = 1/ 7 1+

1
k

Unless the user supplies a value for c, its value is calculated as above. Note that if a different
value is supplied, the resulting distribution will have a mean value which is different from
V .
The annual energy yield is calculated as
cutout

E =Y

P (V ) f (V )dV
cutin

where

P (V ) = power curve, i.e. electrical power as a function of wind speed,


Y = the length of a year, taken as 365 days.
The result is further multiplied by the availability of the turbine, which is assumed for this
purpose to be uncorrelated with wind speed.

78 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

Frequently a steady state power curve is used, combined with a Weibull distribution of hourly
mean wind speeds. For a more accurate calculation, it is desirable to use a dynamically
calculated power curve given as the average power from a series of simulations based on a
model of the turbulent wind field. It is common practice to use 10-minute simulations to
capture the effects of turbine dynamics and wind turbulence. Strictly speaking, the
appropriate Weibull distribution to use in this case would be one representing the distribution
of 10-minute mean wind speeds in a year. This will typically have a slightly smaller shape
factor than that for hourly means.

9.8 Ultimate loads


The ultimate loads calculation, which is often required for certification calculations, is simple
in concept: the results of a load case simulation are analysed to find the times at which each
of a number of specified loads reaches its maximum and minimum values. The simultaneous
values of all the loads at each of those instants is reported.
A further calculation named ultimate load cases further analyses the results of a number of
ultimate loads calculations for different groups of load cases, and generates a histogram
showing the load cases in which the maximum and minimum values of each load occurred
within each group.

9.9 Flicker
The Flicker calculation generates short-term flicker severity values (Pst), either from a voltage
time history, or from time histories of active and reactive power. Such time histories are
available from simulations with the full electrical model of the fixed speed induction
generator, and also with the variable speed generator model.
The flicker severity is a measure of the annoyance created by voltage variations through
perception of the resulting flicker of incandescent lights. The calculation of flicker from a
voltage time history is defined in [9.5]. An algorithm conforming to this standard is
incorporated into the Bladed post-processor. It can also calculate flicker from a time history
of active and reactive power. In this case a voltage time history is calculated first, and this
can be calculated for any given network impedance to which the wind turbine might be
connected. In fact the flicker for several different network impedances can be calculated in a
single calculation. The network impedances are entered as a set of short circuit power levels
and network angles, the network angle being arctan(X/R), where X and R are the network
reactance and resistance respectively. The voltage is calculated as the solution of the
following equation:
U4 + U2(2{QX - PR} - U02) + (QX - PR)2 + (PX + QR)2 = 0
where U0 is the voltage at the infinite busbar, and P and Q are the active and reactive power
respectively.

79 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

10. REFERENCES
2.1

Glauert H, An aerodynamic theory of the airscrew, Reports and memoranda, AE. 43,
No 786, January 1922

2.2

Prandtl L and Tietjens O G, Applied Hydro and Aeromechanics, Dover Publications,


1957

2.3

Pitt D M and Peters D A, Theoretical prediction of dynamic inflow derivatives,


Vertica, Vol. 5, No 1, 1981

2.4

Gaonkar G H, Sastry VV,, Reddy T S R, Nagabhushanam J and Peters D A, The use


of actuator disc dynamic inflow for helicopter flap lag stability, 8th European
Rotorcraft Forum, France, Sept. 1982

2.5

Tuckerman L B, Inertia factors of ellipsoids for use in airship design, NACA Report
No 210, 1925

2.6

Rasmussen F R, Petersen S M, Larsen G, Kretz A and Andersen P D, Investigations


of aerodynamics, structural dynamics and fatigue on Danwin 180 kW, Ris M-2727,
June 1988

2.7

Snel H, Houwink R, Bosschers J, Piers W J and van Bussel G J W, Sectional


prediction of 3D effects for stalled flow on rotating blades and comparison with
measurements, EWEC 93, Travemunde, March 1993

2.8

Leishman J G and Beddoes T S, A semi-empirical model for dynamic stall, Journal


of the American Helicopter Society, July 1989

2.9

Harris A, The role of unsteady aerodynamics in vertical axis wind turbines, Recent
developments in the aerodynamics of wind turbines, BWEA workshop, University of
Nottingham, February 1990

2.10 Galbraith R A McD, Niven A J and Coton F N, Aspects of unsteady aerodynamics of


wind turbines, Recent developments in the aerodynamics of wind turbines, BWEA
workshop, University of Nottingham, February 1990
4.1

Ahmed-Zaid S and Taleb M, Structural modelling of small and large induction


generators using integral manifolds, IEEE trans. Energy Conversion 6, 3, September
1991.

4.2

Park R H, Two-reaction theory of synchronous machines, Trans AIEE 48, 1929.

4.3

Bossanyi E A, Investigation of torque control using a variable slip induction generator,


ETSU WN 6018, ETSU, 1991.

4.4

Pedersen T K, Semi-variable speed operation - a compromise? Wind Energy


Conversion 1995, 17th BWEA Conference (Warwick), Mechanical Engineering
Publications Ltd.

80 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

7.1

IEC 1400-1, Wind turbine generator systems - Part 1: Safety requirements, First
edition, 1994-12.

7.2

Powles S R J, The effects of tower shadow on the dynamics of a HAWT, Wind


Engineering, 7, 1, 1983.

7.3

Veers P S, Three dimensional wind simulation, SAND88 - 0152, Sandia National


Laboratories, March 1988.

7.4

Engineering Sciences Data Unit, Characteristics of atmospheric turbulence near the


ground. Part II: Single point data for strong winds, ESDU 74031, 1974.

7.5

Engineering Sciences Data Unit, Characteristics of atmospheric turbulence near the


ground. Part II: Single point data for strong winds (neutral atmosphere), ESDU 85020,
1985 (amended 1993).

7.6

Engineering Sciences Data Unit, Characteristics of atmospheric turbulence near the


ground. Part III: Variations in space and time for strong winds (neutral atmosphere),
ESDU 86010, 1986 (amended 1991).

7.7

IEC 1400-1, Wind turbine generator systems - Part 1: Safety requirements, Second
edition, 1997.

7.8

Ainslie J F, Development of an eddy viscosity model for wind turbine wakes,


Proceedings of 7th BWEA Wind Energy Conference, Oxford 1985.

7.9

Ainslie J F, Development of an Eddy Viscosity model of a Wind Turbine Wake,


CERL Memorandum TPRD/L/AP/0081/M83, 1983.

7.10 H Tennekes and J Lumley, A first course in turbulence, MIT Press, 1980.
7.11 L Prandtl, Bemerkungen zur Theorie der freien Turbulenz, ZAMM, 22(5), 1942.
7.12 Ainslie J F, Calculating the flowfield in the wake of wind turbines, Journal of Wind
Engineering and Industrial Aerodynamics, Vol 27, 1988.
7.13 Taylor G J, Wake Measurements on the Nibe Wind Turbines in Denmark, National
Power, ETSU WN 5020, 1990.
7.14 Quarton D C and Ainslie J F, Turbulence in Wind Turbine Wakes, Wind
Engineering, Vol. 14 No. 1, 1990.
7.15 U Hassan, A Wind Tunnel Investigation of the Wake Structure within Small Wind
Turbine Farms, Department of Energy, E/5A/CON/5113/1890, 1992.
7.16 Vermeulen P and Builtjes P, Mathematical Modelling of Wake Interaction in Wind
Turbine Arrays, Part 1, report TNO 81-01473, 1981.
7.17 Vermeulen P and Vijge J, Mathematical Modelling of Wake Interaction in Wind
Turbine Arrays, Part2, report TNO 81-02834, 1981.

81 of 82

Garrad Hassan and Partners Ltd

Document: 282/BR/009

ISSUE:011

FINAL

8.1

Goda Y, A Review on Statistical Interpolation of Wave Data, Report of the Port and
Harbour Research Institute, Vol. 18, No. 1, March 1979.

8.2

Hogben N and Standing R, Experience in Computing Wave Loads on Large Bodies,


OTC 2189, Offshore Technology Conference, Houston, 1975.

8.3

Wheeler J D, Method for Calculating Forces Produced by Irregular Waves, J. Petr.


Techn., pp.359-367, March 1970.

8.4

Gudmestad O T, Measured and Predicted Deep Water Wave Kinematics in Regular


and Irregular Seas, Marine Structures, Vol. 6, pp.1-73, 1993.

8.5

Chaplin J R, The Computation of Non-Linear Waves on a Current of Arbitrary NonUniform Profile, Den Report OTH 90 326, HMSO, 1990.

8.6

Dean R G, Stream Function Representation of Nonlinear Ocean Waves, Journal of


Geophysical Research, Vol.70, No. 18, Sept. 1965.

8.7

Dean R G, Stream Function Wave Theory: Validity and Application, Proceedings of


the Santa Barbara Specialty Conference, Ch. 12, Oct. 1965.

8.8

Dalrymple R A, A Finite Amplitude Wave on a Linear Shear Current, Journal of


Geophysical Research, Vol. 79, pp. 4498-4504, 1974.

8.9

Dalrymple R A, A Numerical Model for Periodic Finite Amplitude Waves on a


Rotational Fluid, Journal of Computational Physics, Vol. 24, pp. 29-42, 1977.

9.1

Rice S O, Mathematical analysis of random noise, Selected papers on noise and


stochastic processes, ed. N Wax, 1959.

9.2

Davenport A G, Note on the distribution of the largest value of a random function


with application to gust loading, Proc. Inst. Civil Eng. 28, pp187-196, 1964.

9.3

Madsen P H, Frandsen S, Holley W E, Hansen J C, Dynamics and fatigue damage of


wind turbine rotors during steady operation, Ris report R-512, 1984.

9.4

Miner M A, Cumulative damage in fatigue, Transactions of the American Society of


mechanical Engineers, Vol. 67, A159-A164, 1945.

9.5

Flickermeter functional and design specification, BSEN60868, 1993, and evaluation of


flicker severity, BSEN60868-0, 1993, equivalent to IEC 868-0, 1991.

82 of 82

You might also like