You are on page 1of 117

Bladed Theory Manual

Version 4.3

DISCLAIMER

Acceptance of this document by the client is on the basis that Garrad Hassan
& Partners Ltd is 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.

COPYRIGHT

All rights reserved. Duplications of this document in any form are not allowed
unless agreed in writing by Garrad Hassan & Partners Ltd.

© 2012 Garrad Hassan & Partners Ltd.

Garrad Hassan & Partners Ltd.


St Vincent’s Works, Silverthorne Lane, Bristol BS2 0QD England

www.gl-garradhassan.com
1. INTRODUCTION

1.1 Purpose

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:


 Aerodynamic and structural properties
 Power train and electrical systems
 Sensors and actuators
 Control and safety systems

 Specification of environmental inputs and load cases:


 Wind field
 Waves and currents
 Earthquakes
 Turbine faults
 Electrical network disturbances

 Rapid calculation of steady-state performance characteristics, including:


 Aerodynamic information
 Performance coefficients
 Power curves
 Steady operating loads
 Steady parked loads

 Dynamic simulations covering all turbine states:


 Normal running
 Start-up
 Normal and emergency shut-downs
 Idling
 Parked

 Post-processing of results to obtain:


 Basic statistics
 Periodic component analysis
 Probability density, revs at level, peak value and level crossing analysis
 Spectral analysis
 Cross-spectrum, coherence and transfer function analysis
 Rainflow cycle counting and fatigue analysis
 Combinations of variables
 Dynamic power curve and annual energy yield
 Ultimate loads (identification of worst cases)
 Statistical extrapolation of extreme loads
 Flicker severity
 Presentation: results may be presented graphically and can be combined into a word
processor compatible report.

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.

Many of the models and theoretical methods incorporated in Bladed have been extensively
validated for wind turbines 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
The code has also been compared with measurements from a scaled tidal turbine device at
Southampton University.

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

1.3 Support

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: +44 (0)117 972 9900


Fax: +44 (0)117 972 9901
E-mail bladed@gl-garradhassan.com

1.4 Documentation

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

1.5 Acknowledgements

Bladed was developed with assistance from the Commission of the European Communities
under the JOULE II programme, project no. JOU2-CT92-0198. .
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 flow 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
fluid behind it experiencing a reduction in pressure compared to free stream conditions. As the
fluid proceeds downstream the pressure climbs back to the free stream value resulting in a
further slowing down of the flow. There is therefore a reduction in the kinetic energy in the flow,
some of which is converted into useful energy by the turbine.

In the actuator disk model of the process described above, the flow velocity at the rotor disk Ud is
related to the upstream wind velocity Uo as follows:

U d  (1  a )U o

The reduced flow velocity at the rotor disk is clearly determined by the magnitude of a, the axial
flow induction factor or inflow factor.

By applying Bernoulli’s 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) 2

where  is the fluid density and A the area of the rotor disk.

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

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)
1 16
The maximum value of the power coefficient CP occurs when a is /3 and is equal to /27 which
is known as the Betz limit.
1
The thrust coefficient CT has a maximum value of 1 when a is /2.

2.1.2 Wake rotation


The actuator disk concept used above allows an estimate of the energy extracted from the flow
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 flow 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 

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 W 2 (CL cos  CD sin  )cdr


2

where W is the magnitude of the apparent flow speed vector at the blade element,  is
known as the inflow angle and defines the direction of the apparent flow 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 V 2 S )
2
and

CD  D /( 1 V 2 S )
2
where L and D are the lift and drag forces, S is the planform area of the aerofoil and V is the
flow 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 W 2 r (CL sin   CD cos )cdr


2
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 ,  g 2 /(1  g 2 )

where

Bc (CL cos  CD sin  )


g1  H
2r 4 F sin 2 

and

Bc (CL sin   CD cos )


g2 
2r 4F sin  cos

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  1.0

4a(1  a)
for a  0.3539,H 
(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 1  tol

and

ak'  ak' 1  tol

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

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.

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 a r  and


not the induction factors a and a which are frozen in time.

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 flow 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.
For a blade element, bounded by radii R1 and R2 , and subject to uniform axial flow at a flow
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 R 3
3
Therefore the thrust coefficient associated with the annulus can be derived to give:

16 ( R23  R13 )
CT  4a(1  a)  a
3U o ( R22  R12 )

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.

2.3 Steady stall

The representation and to some extent the general understanding of aerodynamic stall on a
rotating 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 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 foil 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.

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 also calculated according to the Beddoes-
Leishman model.
Modifications to the Beddoes-Leishman model are as follows:
 Modification of the chordwise force expression to give agreement with steady aerofoil
data.
 The coefficients for the circulatory part of the indicial response are given by Jones’
approximation to the Wagner function that models a thin airfoil undergoing a step
change in angle of attack in incompressible flow: A1 = 0.165, A2 = 0.335, b1 = 0.0455
and b2 = 0.3.
 Addition of vortex component in the chordwise direction according to the results of
[2.11].
 Attached flow contributions applied only to the proportion of the aerofoil for which flow is
attached.
 Expression for circulatory part of pitching moment coefficient based on assumption of
linear part and an additional part due to stall. These are weighted by the time-lagged
separation position so that in limiting steady conditions the quasi-steady pitching
moment is recovered.
 An additional term modelling the contribution to the pitching moment from drag is
included. This is based on the variation of windspeed over the aerofoil chord length
when the aerofoil is pitching and is important for the torsional stability of heavily stalled
aerofoils.
3. STRUCTURAL DYNAMICS

In the early days of the industry, wind turbine design was undertaken on the basis of quasi-
static aerodynamic calculations, while the effects of structural dynamics were either ignored
completely or included through the use of estimated dynamic magnification factors. From the
late 1970’s 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 software packages for
solving problems of structural dynamics is challenging in the case of wind turbines. This is
because of the presence of rigid body motion of one structural component, i.e. the rotor, with
respect to another, i.e. the tower or another support structure type. In principle, the standard
finite element method only considers structures in which the deflection occurs about an initial
reference position, and for this reason the finite element models that have been developed for
wind turbine in the past have been tailored to deal with this problem.

Dynamic wind turbine models commonly used as the basis of design calculations involve a
modal representation of the deformation state. 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. Another important advantage is that the
structural damping of flexible components can be described conveniently and realistically as
modal damping. Obviously the number and type of modal degrees of freedom used to
represent the dynamics of a particular wind turbine depends on the configuration and
structural properties of the machine.

In principle, a wind turbine structure may be considered as a collection of a set of


interconnected structural components that may undergo large rotations relative to
neighbouring components, but also relative to an inertial coordinate system. The natural
choice for modelling structural dynamics of said mechanical systems is the multi-body
dynamics approach that emerged in the late 1980’s, initially for rigid components or bodies
[3.1], but later also for flexible components [3.2-3]. The multi-body dynamics approach is now
used as an integrated part of the design process in the automotive and the aircraft industry,
but it has also been used extensively in the space industry. The structural model in Bladed is
based on this approach combined with a modal representation of the flexible components like
the blades and the tower. This ensures a consistent formulation of the dynamic equations and
facilitates the modelling of the turbines based on new structural concepts.

3.1 The multi-body dynamics approach

The multi-body dynamics approach used for the Bladed structural model was originally
proposed by Dr. J.P. Meijaard, Nottingham University, (presently University of Twente, in the
Netherlands) under commission of Garrad Hassan and Partners Ltd [3.4], and it was
developed particularly for modelling wind turbine structural dynamics. The approach assumes
a tree-like structure, which means that structural loops cannot be described.

In general, the structural components may be assumed to be rigid, such as yaw and blade
bearings, or flexible such as support structures, towers and blades. While rigid components
are relatively easy to model, flexible components are more complicated as the motion of a
material point of this type of components is generally caused by rigid body motion combined
with relative motion due to the deformation. A description of the applied method for modelling
flexible components is given in Section 3.2.
The rigid body motion of a component is described in terms of the motion of a set of local
component nodes that are characteristic material points, where the motion of the component
is assumed to be known. Components can only be interconnected at the nodes, and a
connection is imposed by the usual finite element technique by linking the nodes of the
components involved in the connection to a global structural node. Due to the assumption of
the tree-like structure it is convenient to subdivide the nodes of a component into a proximal
node that is linked to a node of its predecessor, and distal nodes that may be linked to nodes
of successors. For example a yaw bearing has two nodes, i.e. one proximal node attached to
the tower and one distal node attached to the main frame.

For all components a local Cartesian coordinate system is attached to the proximal node such
that the position of origin and the orientation are defined by the position and orientation,
respectively, of this node. This local coordinate system is mainly used for flexible components
that are described in more details in Section 3.2.

The deformation state of a component is described by generalised strains that represent the
degrees of freedom associated with the component. For example a yaw bearing has one
generalised strain, which represent the yaw angle. The approach used also allows prescribed
strains, which are particularly important in the case of stick-slip friction in bearings, where the
bearing may stick if the absolute value of the angular velocity (i.e. the strain rate) approaches
zero. It is noted that prescribing a strain element will reduce the effective order of the system
of equilibrium equations by one.

In general, the relative motion of the distal nodes is constrained, for which reason the position
and orientation of a distal node are expressed as functions of the position and orientation of
the proximal node and the generalised strains. From this fundamental transformation it is
straightforward to derive the corresponding transformations for the velocity and the
acceleration. The constraints are conveniently expressed in terms of two constraint matrices
relating to the nodal velocities and the strain rates. In general, the constraint matrices are
time-dependent functions of the position and orientation of the proximal node as well as the
strains.

Components may have mass or may be considered massless. The generalised inertia forces
are derived from the principle of virtual work, where the inertia force density is expressed
according to D’Alambert’s principle. The material velocity and acceleration are derived from a
fundamental displacement hypothesis that defines the absolute displacement of all material
points of the component as a function of the relative position, the nodal position and
orientation and the strain. The result of this analysis shows that the inertia force can be
expressed in terms of a mass matrix multiplied by the acceleration vector plus a vector of non-
linear inertia forces and stresses, i.e. centrifugal and Coriolis forces.

In principle, external loads can only be applied as nodal loads due to the nodal velocities or
generalised stresses due to the generalised strain rates. For distributed loads like wind and
wave loads the corresponding applied nodal loads and generalised stresses are calculated
according to the principle of virtual work. For a yaw bearing the applied generalised stress is
simply the moment applied by the yaw actuator.

Gravity loads are conveniently considered as a distributed applied body forces.

The resulting equilibrium equations of a component are derived by collecting all generalised
forces acting on the component. The effect of the distal node(s) constraints is described by
Lagrange’s methods in terms of internal forces that are expressed by the constraint matrices
and a set of yet unknown Lagrange multipliers [3.5]. A detailed analysis shows that the
resulting component equilibrium equations and the transformation for the acceleration may be
expressed in matrix form as
 M crr M cr  Dcr T   v c   0   0   fac  f0c   fic 
 M c T M c D c T   ε c    C c  εc   K c  εc   σ c  σ c    σ c 
 rc  
          a c 0  i 
 Dr D 0   λc   0   0   a 2   0 
c

where
v c is a vector of nodal velocities
εc is a vector of generalised and prescribed strains
λc is a vector of Lagrange multipliers corresponding to the constraints
fic and σ ic are vectors of non-linear inertia forces and stresses dual to nodal velocities
and strain rates
fac is a vector of applied nodal forces
σ ca is a vector of applied generalised stresses dual to generalised strain rates
f0c is a vector of joint reactions dual to nodal velocities
σ c0 is a vector of generalised constraint stresses corresponding to prescribed strains
a c2 collects the convective terms (quadratic in nodal velocities) in the transformation
for the acceleration
M crr , M cr and M c are the structural mass matrix partitions dual to nodal velocities and
strain rates
C c is the structural damping matrix dual to strain rates
K c is the structural stiffness matrix dual to strain rates
D cr and D c are the constraint matrix partitions relating to nodal velocities and strain
rates

Obviously it is not possible to solve this equation due to the unknown joint reaction forces f0c
(section forces), which originates from the connection to other components. In order to solve
the system it is necessary to collect all the component equilibrium equations into a global
system of the complete structure, which is done using the standard finite element assembly
process [3.5]. This global system of equations has almost the same form as the component
equations and is written in matrix form as

 M rr M r  DTr   v   0   0   fa   f i 
 M T M DT   ε    C  ε   K  ε  σ  σ    σ 
 r  
          a 0  i
 r
D D  0      
λ 0 0   a 2   0 

The main difference between component equations and the global system of equations is that
the joint reaction forces do not appear in the latter as they have been cancelled out by the
assembly process. Consequently the above system can be solved directly with respect to the
nodal accelerations v , the strain accelerations ε , and the Lagrange multipliers λ .

3.2 Modelling flexible components using the multi-member modal


approach

In the standard wind turbine model in Bladed the blades and the support structure are
modelled as single linear flexible components (bodies) using the modal approach, where the
deformation is represented by a linear combination of some pre-calculated mode shape
functions or simply modes. The scalars of this linear combination are the modal amplitudes
[3.6] that represent the generalised strains and hence the degrees of freedom of the
component. It is important to note that the mode shape functions are constant in time and
varies only spatially.
The mode shape functions for a flexible component are calculated from a fundamental
structural model based on standard linear finite element technique [3.5], which means that the
deformation of a flexible component is generally assumed to be small. The fundamental finite
element model basically assumes that the flexible components are linear space beams or
more general space frames that comprise assemblies of multiple members modelled by
Timoshenko beam elements. It is also noted that the fundamental finite element model is also
used for calculating the internal forces (stress resultants) of the flexible component as
described in Section 3.5.

The applied beam element may be considered as an extension to the standard three-
dimensional engineering Timoshenko beam element [3.7] with two nodes or stations located
at the two ends. This element has twelve fundamental degrees of freedom, i.e. three
translational and three rotational degrees of freedom at both stations. At all intermediate
points the deflection is calculated interpolation functions that are derived from a set of
homogenous equilibrium equations valid for prismatic beam elements. It is important to note
that this beam element includes the effect of shear deformation that may be important for
some support structures, in particular complicated offshore foundations. The magnitude of the
shear deformations relative to bending deformations for an element may be evaluated by the
element shear parameter conveniently defined as

12 EI e
e  2
l e GAe
where
EI e is the bending stiffness
GAe is the corresponding shear stiffness
l e is the element length

The beam element supports an arbitrary spatial variation of the stiffness along the beam
element, but in the present implementation it is assumed that the bending stiffness and
2
torsional rigidity vary with H tw, where H is the cross-section structural height and tw is the
cross-section wall thickness that both are interpolated linearly over the element, whereas the
axial stiffness and the shear stiffness are assumed to vary linearly. The orientation of the
element is defined by the position of the two ends as well as the orientation of the principal
axis around the neutral axis (elastic centre). The effect of possible coupling between bending
and torsion is included in terms of the position of the shear centre (torsion centre), and a
transformation between displacements and forces relating to the shear centre and the neutral
axis is included. The resulting stiffness matrix is constant and calculated by numerical
integration. For prismatic elements, where the shear centre is located at the neutral axis, the
stiffness matrix is identical to the standard engineering Timoshenko beam element [3.7].

An important feature of the derived method is that some fundamental degrees of freedom
may be constrained, which is particularly useful in cases where the effect of elongation and/or
torsion can be neglected. In order to enable the description of rigid connection the deflection
of a beam element may also be constrained completely. The constraints are modelled in
terms of a constant constraint matrix together with Lagrange’s method [3.5].

Second-order effects of the axial forces are included in terms of a geometric stiffness matrix
(stress stiffening) that is calculated from the derivatives of the interpolation functions [3.6]. For
the blades the dominating part of the axial force is caused by centrifugal forces for which
reason the term centrifugal stiffness is traditionally used in this case.

The selection and calculation of mode shape functions follows the idea that was originally
suggested by Craig [3.8] as a modification of the celebrated Craig-Bampton method [3.9]. For
both methods the stations are subdivided into boundary stations that may couple to other
components and interior stations that do not couple. The boundary stations also represent the
component nodes that may link to nodes of other components. In particular the station
representing the proximal node is constrained completely in order to exclude rigid body
displacement modes. With the applied method the modes are generally selected as the union
between attachment modes that may couple to other components and normal modes that
may be considered as internal vibration modes. The attachment modes are calculated from
the component stiffness matrix by a static equilibrium, where the component is fixed at the
proximal node and point loads are applied in turn at the distal nodes. The normal modes are
calculated from the mass and stiffness matrices by a generalised eigenvalue problem, where
the component is fixed at all boundary stations. The frequencies of the attachment modes are
calculated by Rayleigh’s method [3.6], while the frequencies of the normal modes result from
the eigenvalue problem. These frequencies are solely used for describing damping as
explained below.

The modes are orthogonalised in a particular form that ensures that the stiffness matrix
relating to the generalised strains is diagonal.

Inertia forces acting on the element and the proximal node are derived as described in
Section 3.1 from the fundamental displacement hypothesis using the principle of virtual work.
An important feature of the derived method is that the inertia forces are expressed directly in
terms of the modal amplitudes, i.e. the strains and corresponding derivatives as originally
suggested by Shabana [3.2]. In principle this means that the time for calculating the
accelerations scales with the number of mode shapes rather than the number of beam
elements of the flexible component.

Structural damping is modelled as modal damping [3.6] in terms of a set of damping


coefficients (damping ratio) that relate to the mode shape functions. These coefficients are
defined as input parameters for the model and may usually be measured directly, for example
by exiting a mode and measure the decay of the succeeding oscillation.

3.2.1 Blade modes


The motion of the tapered and twisted rotor blades is among the most complex phenomena
related to the structural dynamics of a wind turbine. For the standard blade model in Bladed,
a blade is represented by one component and has a proximal node at the root flange. As
described in Section 3.2 this implies that the mode shape functions are represented by
normal modes only, which is the classical approach for selecting the vibration modes of a
wind turbine blade. Each mode is defined in terms of the following parameters:

 Modal frequency, i
 Modal damping coefficient, i
 Mode shape represented by a vector of the displacement at the stations

where the subscript i indicates properties related to the i’th mode.

The modal frequencies and mode shapes of the blade are calculated based on the following
parameters that generally are defined at each radial station:

The position of neutral axis


The position of the neutral axis is defined by the position of the stations defined by the radius
and two transversal coordinates defined in the coordinate system attached to the blade root.
The neutral axis determines the longitudinal shape of the blade, which means that this
parameter also determines pre-curving (pre-bending). It is noted that in the field of aero-
elasticity it is common practise to use the term elastic centre for the neutral axis.

The cross-section position of the mass centre


The cross-section position of the mass centre is given in the flapwise and edgewise
directions.

The orientation of the principal axis of mass


The orientation of the principal axis of mass defines the orientation of the local coordinate
system that is used for defining the mass properties.

The mass distribution along the blade


The mass distribution is defined as the local mass density in addition to the magnitude and
location of any discrete, lumped masses.

The radii of gyration


The radii of gyration are defined in the flapwise and the edgewise directions defined by the
orientation of principal axis of mass. It is noted that the torsional moment of inertia is
determined by the radii of gyration together with the mass distribution.

The structural twist angle along the blade


The structural twist angle defines the orientation of the local element system and hence the
direction of the flapwise and edgewise stiffnesses resolved through the local twist angle.

The cross-section position of the shear centre


The cross-section position of the shear centre defines the origin of the shear forces and
therefore the torsion centre. This parameter is only needed if the shear stiffness or the torsion
rigidity is defined.

The bending stiffness along the blade.


The bending stiffness is defined in local flapwise and edgewise directions defined by the
structural twist.

The shear stiffness along the blade


Like the bending stiffness the shear stiffness is defined in local flapwise and edgewise
directions. Usually shear deformations for the blades may be neglected in which case the
shear stiffness is not needed.

The torsional rigidity along the blade


The torsional rigidity is defined with respect to the shear centre. Usually the elongation of the
blade may be neglected in which case the axial stiffness is not needed.

The frequencies and mode shapes of the rotor modes are computed from the eigenvalues
and eigenvectors of the fundamental finite element model of the blade structure. The finite
element model of the blade is based on the use of three-dimensional beam elements to
describe the mass and stiffness properties of the rotor blades as described in Section 3.2.

The outputs from the modal analysis of a blade are the modal frequencies and mode shapes
defined by a vector of the displacements of all stations, i.e. three translational components
and three rotational components. The modal damping coefficients are an input defined by the
user and may be used to represent structural damping.

3.2.2 Tower modes (axisymmetric model)


The standard axisymmetric tower model in Bladed has one proximal node at the base and
one distal node at the top. This implies that the mode shape functions are represented by a
combination of attachment modes and normal modes, which represent the deflection in the
fore-aft and side-side directions as well as the torsion and elongation. As for the rotor, the
tower modes are defined in terms of their modal frequency, modal damping and mode shape.

The calculation of modal frequencies and mode shapes of the tower make use of the
following parameters:

The position of neutral axis


Basically the axisymmetric tower is assumed to be defined along a straight line, which means
the neutral axis is straight. The stations defining the neutral axis are therefore solely defined
by the height.
The mass distribution along the tower
The mass distribution is defined as the local mass density at each tower station height in
addition to the magnitude and location of any discrete, lumped masses.

The bending and shear stiffness along the tower


The tower is assumed to be axisymmetric with the bending and shear stiffness therefore
independent of bending direction.
The mass, inertia and stiffness properties of the 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 stiffness.

The mass and inertia of the nacelle and rotor


For the 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. As the normal
modes do not couple to other component it appears that only the frequency of the attachment
modes are affected. This means that the mass and inertia of the nacelle and rotor only affect
the resulting damping of the attachment modes.

The frequencies and mode shapes of the tower modes are computed from the eigenvalues
and eigenvectors 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 representing fore-aft and side-side deflection and possibly torsion. The modal
damping coefficients are an input defined by the user and may be used to represent structural
damping.

3.2.3 Tower modes (multi-member model)


Once again a modal representation is used, but in this case an arbitrary structure consisting
of any number of straight interconnected members is permitted. Since the tower is not
assumed to be axisymmetric, the modes are generally three-dimensional and contain all six
degrees of freedom at each node, and there may be a foundation at more than one node.

3.2.4 Coupling between rotor and tower modes


The modal analysis calculation generates the component modes of vibration for the blades
and the tower . These are linked together in the Bladed equations of motion to give coupled
system modes. The frequencies of the coupled system modes will not be the same as those
of the uncoupled component modes; the coupled mode frequencies can be found using the
Campbell diagram calculation.

3.3 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 relatively complicated, and the following only gives a
brief description of the theory.

3.3.1 Degrees of freedom


The degrees of freedom involved in the equations of motion for the structural dynamic model
for Bladed are as follows:

 Blade deflection
 Nacelle yaw
 Tower fore-aft, side-side and torsional deflection (axisymmetric tower model)
 General tower deflection (multi-member tower model): a large number of modes is
allowed, including the torsional and axial degrees of freedom.

In addition, the following dynamics can also be included as required:

 A sophisticated representation of the power train dynamics as described in Section 4.1 of


this manual.
 A range of different representations of generator and power converter dynamics, including
both mechanical models and electrical models which can include network interactions
(Section 4.2).
 A range of pitch actuator models, from simple passive models to detailed representations
of electric servo drives and hydraulic actuators (Section 5.6).
 Teeter restraints, passive blade vibration dampers and tower dampers, and yaw system
dynamic response.
 Transducer dynamics for control signals.
 All controller dynamics.

3.3.2 Formulation of equations of motion


As described briefly in Section 3.1 the equations of motions of the complete system have
been derived using the multi-body dynamics approach based on the principle of virtual work.
It appears that the solution of the resulting equations mentioned in Section 3.1 is generally
difficult to obtain as the augmented mass matrix including the constraint matrices are
generally ill-conditioned. The system is therefore transformed into a system where the only
unknown is the strain accelerations ε using the so-called constraint elimination method [3.3]
together with the transformation for the velocity given in the form Dcr v  Dc ε  0 . The final result
of this straightforward transformation can be written in the conventional form

Mε  Cε  Kε  σ

In cases with no prescribed strains it is straightforward to show, that the three system
matrices appearing on the left-hand side of the above equation become

M  M  DTr MrrDr  DTr Mr  MTr Dr , C  C , K  K 

where D r  Dr1D is the time-dependent part of the reduction matrix. The right-hand side
stress vector of the global system of equations becomes

σ  σa  σ0  σi  MTr  g 2  DT r  fa  fi  M rr g 2 

where g2  Dr1a2 .

In general, the system mass matrix M is full, due to the coupling of the degrees of freedom,
and it contains periodic coefficients because of the time-dependent interaction of the
dynamics of the rotor and tower. The system damping and stiffness matrices C and K are
generally diagonal and constant.

Because of their complexity, further details of the equations of motion are not presented here.

3.3.3 Solution of the equations of motion


The equations of motion are solved by time-marching integration of the ordinary differential
equations using a variable step size, fourth order Runge-Kutta integrator [3.10].
3.4 Calculation of structural loads

The structural loads at the nodes, e.g. the blade root, the power train and the tower base are
calculated as the joint reactions by the component equilibrium equation described in Section
3.1. The inertial loads are calculated by integration of the mass properties and the total
acceleration vector at each station according to the principle of virtual work. The total
acceleration vector includes modal, centrifugal, Coriolis and gravitational components.
Second-order effects of axial forces due to structural deflections are taken into account, i.e.
the contribution to bending moments caused by the applied axial forces acting on the
deflected structure. For example, the contribution to tower base bending moment caused by
the weight of the nacelle takes into account the true position of the nacelle centre of gravity
including the deflection of the tower top.

3.5 Calculation of structural deflections and internal forces of flexible


components

With the modal model, the deformed shape of the flexible components like the blades and
tower at any instant is a linear combination of the selected mode shape functions. With a
reduced number of modes, the resulting deformation may therefore not be accurately
predicted, which means that it is not possible to calculate the internal forces directly from the
deformations as done by standard finite element technique.

In order to calculate the internal forces of flexible components the deformation at all station is
therefore calculated from a static equilibrium equation, where the applied force is calculated
as the sum of all external forces including the inertia forces. In case that some fundamental
degrees of freedom are constrained the system is solved with respect to a reduced set of
independent degrees of freedom, and the Lagrange multiplies corresponding to the
constraints are calculated. Finally the internal forces of all beam elements at both ends are
calculated from the fundamental equilibrium equation of the beam element.
nd
In order account for the 2 order effects of the external loads on the calculated section
forces, the external loads are transformed into the non-deformed state, where the static
equilibrium equation is solved, after which the calculated section forces are transformed back
to the deformed state. However, the transformation of the external loads are only done for the
equivalent loads at the stations, so this transformation is not completely correct for distributed
loads like gravity and inertial forces.

3.6 Multi-blade co-ordinate transformation for linearised models


For linearisation calculations or Campbell diagrams it is recommended to select the multi-
blade coordinate transformation, which generates coupled modes referred to the non-rotating
co-ordinate system including the “backward and forward whirling” modes of the rotor. This is
based on theory developed in [3.11] and [3.12]. The linearised model is significantly azimuth-
dependent, but when transformed to non-rotating co-ordinates the resulting model of the
structural dynamics should be only weakly azimuth-dependent. However, for 2-bladed
turbines there is still a strong azimuth dependency.

The basic transformation is as follows:

q
1
q0  b
N b 1
N

q
2
q cn  b cos(n b )
N b 1
N

q
2
q sn  b sin( n b )
N b 1
N

q
1
qd  b ( 1)
b
N b 1

The notations are as follows: qb is a particular degree of freedom for blade b (for example a
flapwise or edgewise degree of freedom). There are N blades, and b is the azimuth angle
for blade b, and n is an integer which goes from 1 to (N-1)/2 if N is odd, and 1 to (N-2)/2 if N is
even. The transformation produces a collective mode q0, a set of cosine and sine modes qcn
and qsn (just one of each for a three-bladed turbine, and none for two-bladed). There is also a
differential mode qd which exists only for even N.

Such a set of modes exists for each degree of freedom. Here are a few examples:

Two blades, N=2:

For each blade degree of freedom there are two rotating degrees of freedom, q 1 and q2, and
they transform into two non-rotating modes, q0 and qd. If qb represents the first flapwise mode
of blade b for example, qo will be a symmetric out of plane rotor mode and qd will be
antisymmetric (like a teeter mode). If qb is the first edgewise mode, q0 will be the collective in-
plane rotor mode and qd the antisymmetric in-plane rotor mode. There are similar sets for the
second and higher blade modes. Other blade modes (e.g. torsional) would be treated in the
same way.

Three blades, N=3:

For each blade degree of freedom there are three rotating degrees of freedom qb for b = 1 to
N, and they transform into three non-rotating modes, q0, qc1 and qs1. q0 is a collective mode
and the other two are like d and q axis modes.

The original state-space model is

q  Aq  Bu
y  Cq  Du

In non-rotating co-ordinates this can be expressed as:

q NR  A NR q NR  B PNR u
y  C PNR q NR  D PNR u
with
 ) U 1
A NR  ( UA  U
B PNR  UB
C PNR  CU 1
D PNR  D

The rotational transformations are applied only to blade modal states, not to any other
individual-blade states such as pitch positions, rates, actuator internal states, etc. The matrix
U just has unit diagonal elements for rows and columns corresponding the states and state
derivatives which are not transformed. For other rows and columns the elements of U
represent the basic transformation defined above for each group of modes. Note that the
elements connecting states and state derivatives also need to be defined by differentiating the
equations of the basic transformation, bearing in mind that the derivative of the azimuth angle
is equal to the rotor speed (which is assumed constant for this purpose). Model inputs and
outputs are not transformed.
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. 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.

The turbine rotor is accelerated by the torque imbalance between the aerodynamic torque
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 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 types of generator model 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)

In each case there is a choice between a mechanical model and an electrical model. The
electrical model allows interactions with the network to be modelled. This is useful for
calculations of electrical flicker, voltage variations, power and reactive power variations and
power factor, and response to network transients such as voltage and frequency transients,
(which wind turbines are increasingly being required to ride through without shutting down).

4.2.1 Fixed speed induction generator – mechanical model


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(  0 )  Q]

where  is the actual generator speed and 0 is the generator synchronous or no-load speed.

The slip slope is calculated as

Pr
h
r (r  0 )

where r is the generator speed at rated power output Pr , given by r = 0 (1 + S/100) where
S is the rated slip in %, and  is the full load efficiency of the generator.

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.

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  xm2
 (4.2.2.1)
X s Rrs

where Xs = xs+xm = sLs, Xr = xr+xm = sLr, and s is the grid frequency in rad/s.

The 2nd order model represents the generator as a voltage source  behind a transient
2
reactance X’ = Xs - xm /Xr, ignoring stator flux transients:

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 ir  js r (4.2.2.2)
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  (4.2.2.3)
Xr
to give

 r  jX s  X  X
T0   s  jssT0   j s vs (4.2.2.4)
 rs  jX   rs  jX 

where
Xr
T0  . (4.2.2.5)
 s rr

The 4th order model is a full d-q (direct and quadrature) axis representation of the generator
which uses Park’s 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

Ls Lr  L  

d is   Lr rs  js Ls Lr  sLm
2
 Lm rr  js Lr Lm (1  s) is   Lr 
 
2

dt ir   Lm rs  js Ls Lm (1  s)
m
  vs
 Ls rr  js L2m  sLs Lr )ir   Lm 
(4.2.2.6)

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 – mechanical model


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:

Qd
Qg 
(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

Qd Num( s)
Qg   g
(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 speed generator – Synchronous generator with fully rated


converter

Notation:
C DC link capacitance, [F]
d, q d, q rotor reference frame
ids , iqs Stator currents in d and q-axis, [A]
ids  ref , iqs  ref Reference stator currents in d and q-axis, [A]
idwt , iqwt Grid side converter d and q-axis currents, [A]
ikd , ikq Damper winding currents in d and q-axis, [A]
iconmax Grid side converter maximum current, [A]
Llkd , Llkq Damper winding leakage inductance in d and q-axis, [H]
Lmd , Lmq Mutual inductance in d and q-axis, [H]
Lls Stator leakage inductance, [H]
X lkd , X lkq Damper winding leakage reactance in d and q-axis, [ohm]
X md , X mq Mutual reactance in d and q-axis, [ohm]
X ls Stator leakage reactance, [ohm]
P Number of pole pairs, [-]
Rs , Rkd , Rkq Stator, d-axis damper, q-axis damper resistance, [ohm]
Rchop Chopper protection resistor, [ohm]
Tc Torque demand from controller, [Nm]
vas , vbs , vcs Stator phase voltages [V]
vds , vqs Stator voltages in d and q-axis, [V]
vdc Wind turbine DC link voltage, [V]
vdc max Wind turbine DC link maximum voltage, [V]
vdc th Wind turbine threshold voltage, [V]
vdctol DC link voltage tolerance for chopper protection, []
vTerm Wind turbine terminal voltage, [V]
vTermmin Wind turbine DC link maximum voltage, [V]
vTermth Wind turbine threshold voltage, [V]
d ,  q Generator controller d and q-axis time constants, [s]

 pll Grid side controller PLL time constant, [s]

 prot Protection filter time constant, [s]

s , r , b Synchronous speed, rotor speed and base speed [rad/s]


ds , qs Stator flux in d and q-axis, [V/(rad/s)]
kd , kq Rotor d and q-axis damper winding flux linkage, [V/(rad/s)]
M , M Rotor flux linkage due to permanent magnet, [V/(rad/s)] & [V]

An electrical model for synchronous generator with fully rated converter is available, as shown
in Figure 4.2. Two four quadrant IGBT converters are used as generator side and grid side
converters. Back-to-back converters connected through a DC link decouple the operating
frequency of the generator from the grid. Therefore the generator can be operated at a
different speed from the grid synchronous speed.

IGBT generator IGBT grid side Grid


side converter converter

Synchronous DC link
generator

Generator Grid side


controller controller

Figure 4.2: Schematic: synchronous generator with fully rated IGBT-IGBT converters

The generator side converter rectifies variable frequency AC power to DC and then the grid
side converter inverts back to grid frequency AC. The rating of the converters has to be higher
than the rating of the generator as the total generator power flows through the converters.
Even though a gearbox is shown in Figure 4.2, a multi-pole synchronous generator can be
used for a direct drive wind turbine.

The electrical system in Figure 4.2, has two main control loops, a generator controller which
manipulates the generator torque, and a grid side controller which maintains the DC link
voltage. Whenever necessary, reactive power can be supplied to the grid by the grid side
converter.

Generator modelling (permanent magnet synchronous generator)


The reference frame used for a synchronous generator is generally fixed to the rotor (rotor
reference frame). The rotor reference frame avoids time varying inductances due to the
salient effect [4.7]. Synchronizing the reference frame to the rotor also enables
implementation of vector control quite simply [4.6]. The d-axis of the reference frame is
0
aligned with the rotor flux vector and the q-axis is 90 ahead of the d-axis. Two short circuited
damper windings are included in each axis of the rotor if there is any damping effect from the
rotor.
Stator differential equations in the rotor reference frame are given in [4.7]

d
vds  Rs ids  r qs  ds
dt
d
vqs  Rs iqs  r ds  qs
dt
The equations of the short circuited damper windings are written as

d
0  Rkd ikd  kd
dt
d
0  Rkqikq  kq
dt
Flux linkages in each circuit are written as

ds  Lds ids  Lmd ikd  M


kd  Lkd ikd  Lmd ids  M
qs  Lqsiqs  Lmqikq
kq  Lkqikq  Lmqiqs

where the permanent magnet produces a rotor flux, M . The salient effect of the rotor is
taken into account by considering different values of the mutual inductance Lmd and Lmq . A
synchronous generator can be represented as (i) a steady state, (ii) a reduced order or (iii) a
full order model [4.7]. The steady state model neglects both stator and rotor transients. The
reduced order model neglects the stator transients while including rotor transients. The full
order model includes both rotor and stator transients.

Unbalanced generator faults


The simulation of unbalanced operation of a synchronous machine requires manipulation of a
large number of equations. This complexity is due to the asymmetry of the machine rotor.
Dynamic equations given in [4.8] and [4.9] are used in Bladed and important equations that
describe different types of generator faults are given below. It is assumed that in the event of
a generator fault, the converter is disconnected from the generator.

Phase-to-phase fault
Assume that phase-b and phase-c are short circuited due an internal generator fault. In order
to have zero phase-a current, the applied voltage to phase-a which will force ias to remain
zero is
1 d a
vas 
b dt
For phase to phase fault,
vbs  vcs

Phase-to-phase to ground fault


This is a very similar situation to the phase-to-phase fault but the short circuited phase-b and
phase-c are now grounded. Therefore vbs  vcs  0

Phase-to-ground fault
During the fault, phase-c is short circuited to the ground and phase-b and phase-a are open.
In order to have zero phase-a and phase-b currents, the following equations need to be
satisfied
1 d a
vas 
b dt
1 d b
vbs 
b dt

Phase-c is grounded, therefore


vcs  0

Generator saturation
Above a certain flux density, the iron core of an electrical machine is subjected to magnetic
saturation. In Bladed, generator saturation is modelled by representing inductance as a
function of current in both d and q-axes. For more information see [4.7].

Demagnetization
Figure 4.3 shows typical demagnetization curves of a permanent magnet machine at different
temperatures. At high temperatures, a demagnetization curve tends to shrink towards the
origin. As this shrinking occurs, the flux available from the magnets reduces. The
demagnetization curve returns to its original shape as temperature drops.

During a generator short circuit, there is a risk of irreversibly demagnetizing the permanent
magnets of the generator. This is because the strong opposing magnetic field created by
short circuit currents forces the magnet to operate in the area of the critical knee. Irreversible
demagnetization permanently damages the torque producing capability of the generator. For
more information on demagnetization see [4.10].

B
Increasing
temperature

Irreversible
demagnetization

Critical knee

Recoil
0 H

Figure 4.3: Typical demagnetization curves of a PM machine

Generator vector controller


A synchronous generator can be controlled using a vector control technique based on rotor
flux orientation [4.6]. The generator electromagnetic torque can be written as

3
Te    P  M  iqs
2
where P is the number of pole pairs of the generator. Therefore for a particular torque
demand ( Tc ) the reference stator current in the q-axis is calculated as

 2  1   1 
iqs  ref        Tc
 3  P    M 

In order to generate maximum torque for a given current, the reference current in the d-axis
( ids ref ) is set to zero for a permanent magnet synchronous generator. Alternatively, the d-
axis current can be used to maintain unity power factor at the generator terminal. When the
required stator voltage magnitude exceeds the converter ceiling voltage, a field weakening
process can be carried out by manipulating the reference current in the d-axis.

The generator side converter functions as a voltage source, therefore the required voltage
vectors are determined using reference and measured currents, PI controllers and decoupling
loops as shown in Figure 4.4 [4.5].

Since maximum AC voltage of a converter is limited by the DC link voltage, it is appropriate to


apply limits on the PI controllers. When a permanent magnet generator operates at top of its
speed range the terminal voltage will be at its maximum. Therefore the PI controller output is
limited to the difference between the DC link voltage and the AC voltage.

r M
  
iqs  ref PI vqs
 

iqs r Lqs

ids r Lds
 
ids  ref  PI vds

Figure 4.4: Control loops of the generator vector controller

Instead of using measured currents for the decoupling loops, the reference current can also
be used for simulation purposes. For simulation purposes the controller can be further
simplified as shown in Figure 4.5. A steady state representation of the vector control scheme
is shown in Figure 4.5 (a). This model neglects all the dynamics associated with the converter
control. Two first order lags are added to the steady state model in Figure 4.5 (b). Time
constants of the first order lags represent any delays in the converter control. Time constant
 q , in q-axis, represents any delay in the torque control loop. The time constant  d is to
represent any delay in the flux control loop.
r M r M
  1  
iqs  ref Rs vqs Rs vqs

iqs  ref 1 qs 
r Lqs r Lqs

r Lds r Lds
 1

ids  ref Rs vds ids  ref Rs vds
 1 d s 
(a) Steady state
(b) First order lag
Figure 4.5: Generator control options

Grid side converter controller


The dq - reference frame used for the grid side converter control rotates at the synchronous
speed of the grid. The q-axis is aligned with the grid voltage and the d-axis is lagging the q-
0
axis by 90 . In order to maintain the DC link voltage, active power coming from the generator
has to be transferred to the grid. Measured DC link voltage is compared with the reference
and the error is fed through a PI controller to obtain the current in the q-axis as shown in
Figure 4.6.

The terminal voltage of the wind turbine is controlled by manipulating the reactive power from
the grid side converter using the current in the d-axis. This controller provides additional
reactive power during a grid fault. In order to ensure that the grid side converter supplies
more reactive current while not exceeding the current limit of the converter, the limit of the
active power current is dynamically adjusted depending on the reactive power current as
shown in Figure 4.6. Instead of controlling the terminal voltage, power factor of the grid side
converter can be controlled. Reference current in d-axis is determined for a given power
factor and active power output.
idwt  max
vTerm  ref  Edwt
1stPId 2ndPId
   

*
idwt
vTerm
X conv
idwt

iqwt
X conv

 max  idwt  max


2 2
icon
  Eqwt
1stPIq 2ndPIq
vdc  ref  *  
 iqwt
vdc

Figure 4.6: Control loops of grid side converter

DC link protection for a grid fault


In the event of a grid fault, the DC link voltage will rise rapidly because the grid side converter
is prevented from transforming all the active power coming from the generator. Therefore, in
order to maintain the DC link voltage below its upper limit, the excess power has to be
dissipated or the generator power has to be reduced.

A chopper resistor on the DC link may be used to dissipate the excess power. The chopper
protection scheme switches a resistance ( Rchop ) ON and OFF at the DC link depending on the
DC link voltage. The controller switches the chopper resistance ON when the DC link voltage
exceeds vdc  max  1  vdc tol  . The chopper resistance is switched OFF when the DC link voltage
drops below vdc  max  1  vdc tol  .

An alternative approach to maintaining the DC link voltage below the upper limit is to reduce
the generator power. The generator power can be reduced rapidly by means of reducing the
generator torque. A droop controller acting either on the terminal or the DC link voltage is
used to reduce the generator torque demand. The droop gain in Figure 4.7 is used as a
multiplier for the torque demand.

Above the terminal threshold voltage or below the DC link threshold voltage, the droop gain is
1. If the droop controller is acting on the terminal voltage, the droop gain is reduced to zero
linearly as the terminal voltage reaches its minimum set point. Similarly, the droop gain drops
towards zero as the DC link voltage reaches its maximum limit (when the droop controller is
acting on the DC link voltage).

r Tc

1
vterm 1

1   prot s
vterm th
vterm  min

vdc 1

vdc  max
vdc th

Figure 4.7: Torque control droop for wind turbine fault ride through

When using the torque control protection method, it may be necessary to drop the generator
torque to zero rapidly as possible to keep the DC link voltage below its limit. This torque step
reduction could induce significant oscillations in the drive train. Therefore the generator torque
can be reduced at a defined ramp rate that is sufficiently slow so as not to excite the shaft-
line. However this method requires a braking chopper to protect the DC link voltage.

4.2.5 Variable speed generator – DFIG electrical model


An electrical model for doubly-fed induction generators is available, as shown in Figure 4.8.
The stator active and reactive power are fed directly to the network, while the rotor active and
reactive power pass through the power converter. The converter efficiency k c results in a loss
of active power.
Grid

DFIG

Crowbar IGBT converters

Torque Voltage or PF
control control

Figure 4.8: Schematic of doubly-fed induction generator system

The converter is controlled by two main control loops, as described in [4.5]: a torque control
loop which works by injecting a quadrature-axis voltage into the rotor circuit, and a voltage or
power factor control loop which works by injecting a direct-axis voltage. Both loops are
essentially PI controllers, with additional terms for decoupling the two loops.

Rotor voltage increases with the slip speed but only until it reaches the rotor side converter
maximum voltage. The voltage limit on maximum speed occurs at minimum grid frequency
and operating with a capacitive power factor. At the top of the operating speed range, the
converter voltage limit will force the generator to absorb a large amount of reactive power
from the grid.

Generator modelling
Generator modelling is similar to the modelling of the synchronous generator given in section
4.2.5. The reference frame used for an induction generator is generally fixed to the stator flux
linkage. The d-axis of the reference frame is aligned with the stator flux vector and the q-axis
0
is 90 ahead of the d-axis.

Stator differential equations are given as

d
vds  Rs ids  s qs  ds
dt
d
vqs  Rs iqs  s ds  qs
dt
The equations of the short circuited damper windings are written as

d
vdr  Rr idr  s  r  qr  dr
dt
d
vqr  Rr iqr  s  r  dr  qr
dt
Flux linkages in each circuit are written as
ds  Lds ids  Lmd idr
qs  Lqsiqs  Lmqiqr
dr  Ldr idr  Lmd ids
qr  Lqr iqr  Lmqiqs

Generator vector controller


A DFIG can be controlled using a vector control technique based on stator flux orientation.
The generator electromagnetic torque can be written as

3  3 v L
Te    P  ds  iqs    P s . m iqr
2  2  s Ls

where P is the number of pole pairs of the generator. Therefore for a particular torque
demand ( Tc ) the reference stator current in the q-axis is calculated as

 2  2  L 
iqr  ref     . s s .Tc
 3  P  vs Lm

It is assumed that the magnetising current is provided the rotor side converter and the
reference magnetising current in the d-axis is calculated as

1
idr  ref  vs
s Lm

The generator side converter functions as a voltage source, therefore the required voltage
vectors are determined using reference and measured currents, PI controllers and decoupling
loops as shown in Figure 4.9. Instead of using measured currents for the decoupling loops,
the reference current can also be used for simulation purposes.

sLm
vs
Ls

 
 vqr
iqr  ref PI
 
 L2 
iqr ss  Lr  m  iqr
 Ls 

 L2 
idr ss  Lr  m  idr
 Ls 

 
idr  ref  PI vdr

Figure 4.9: Control loops of the DFIG vector controller

Grid side converter controller


The grid side converter control is same as the current controller shown in Figure 4.6. The q-
axis is aligned with the grid voltage and maintains DC link voltage. The d-axis current is used
to control the reactive power flow from the grid side converter.

Crowbar protection for a grid fault


During a grid fault the stator voltage drops to a very low value and induces high rotor currents
and this could potentially damage the converter. In order to protect the converter, a crowbar
circuit is connected to the rotor terminals, which is activated in the even of a grid fault to
reduce the impact.

Stator
terminal

Rotor

Converter

Crowbar
Figure 4.10: Crowbar circuit

There are a number of possible design solutions for the crowbar, however for the purposes of
this analysis the crowbar is assumed to be made up of a three phase diode rectifier and a
thyristor as shown in Figure 4.10. During a grid fault, the crowbar thyristor is switched on
when the rotor current reaches its maximum limit.

When the rotor current drops below the maximum limit and there is enough grid voltage the
rotor converter is gradually re-engaged to the rotor circuit. This switchover from the crowbar to
the rotor converter upon clearing the fault turns off the crowbar thyristor and removes the
crowbar from the rotor terminals.

4.2.6 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 continuous-
time PI controller which adjusts the rotor resistance between the defined limits (with 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 4.11.
Torque Current 1 PI with Rotor
demand demand |I| limits resistance

Measured current |I|

Figure 4.11: 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).

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 4.12) 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.

Other turbines
(if required)

Wind
turbine R1 + jX1 R2 + jX2
PCC
Farm Network
interconnection connection
impedance impedance
Infinite
busbar

Figure 4.12: The network model


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,
Figure 5.1: The fixed speed pitch regulated control loop

and the pitch decreases towards the minimum value (usually a negative pitch angle) above
rated.

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 set-
point 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
set-point 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
Figure 5.2: The variable speed stall regulated control loop

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 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
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 =   R Cp() / 2  G
5 3 3

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 where  is the measured generator
2

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
Figure 5.4: Stall regulated variable speed control loops

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.

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

Figure 5.5: The variable speed pitch regulated control loop

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

Figure 5.6: Variable speed pitch regulated operating curve


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

Figure 5.7: Pitch regulated variable speed control loops

remains constant, with the pitch control loop becoming active when the speed exceeds S5.

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

1
y  ( x  y)
T
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 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


1
y  ( x  y)
T
where x is the input and y is the output. The second-order model is represented by

y  2 y   2 ( 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 smoothes 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 term proportional to the axial
and radial forces on the bearing. Sometimes the actuator cannot overcome the applied
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 Measured Pitch rate
Measured
demand from pitch demand from
pitch rate
controller position controller

Bearing
- loads
Ramp control +

Pitching Bearing
moment friction
PID controller

Actuator
torque
limits
Pitch rate
demand
-
+

Pitching
inertia
Acceleration
Passive
limits PID controller
dynamics

Actuator
Passive
torque
dynamics
demand

Passive
dynamics

Actuator
torque

 Pitching
inertia

Pitch rate

Actual pitch
position

Figure 5.8: Pitch actuator options


1.2

0.8
Demand

0.6

0.4 Raw demand


Rate limit
0.2 Acceleration limit
Rate & acceleration limits
0
-0.2 0 0.2 0.4 0.6 0.8 1 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 non-
zero.

Figure 5.9: Ramp control for pitch actuator position demand

Bladed supports both rotary and linear pitch actuators. For linear actuators, for example
using a hydraulic ram, the linkage geometry must be supplied. The principal difference
between the linear and rotary actuator models is that a rotary actuator has a torque limit,
while a linear actuator has a force limit (which translates into a torque limit which varies with
pitch position).

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 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
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
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.

5.7.2 Automatic PI gain calculation


Controller design and tuning is a specialised task. An automatic calculation of PI gains is
possible in Bladed for the torque and pitch controllers of a variable speed pitch feathering
controlled turbine, but the calculation is only valid for a completely rigid turbine in uniform
axial flow, and is therefore only suitable at a very early stage of design, perhaps for optimising
the blade aerodynamic design before taking into account any other details of the turbine.
Otherwise a professional controller design is required.

The equation governing the rotational dynamics for the rigid turbine is:

 Q
J aero  Q gen  Q    Q    Q V V
' ' '

with the following nomenclature:

J Total inertia
Qaero Aerodynamic torque (steady value)
Qgen Generator torque (referred to low speed shaft)
 Rotor speed (deviation from steady value)
 Pitch angle (deviation from steady value)
V Wind speed (deviation from steady value)
Q ' Partial derivative of torque w.r.t. rotor speed
Q ' Partial derivative of torque w.r.t. pitch angle
Q 'V Partial derivative of torque w.r.t. wind speed

For the torque PI controller, the plant can be represented by


Js  Q gen  Q ' 

The plant P is the transfer function from generator torque to rotor speed:

1
P
Q '  Js

 
Ignoring all sensor and actuator dynamics, the PI controller is C  1s K i  sK p where Ki and
Kp and the integral and proportional gains, and the open loop transfer function H = P*C is
given by

K  sK p 
H
i


s Q '  Js 
The closed loop transfer function is H/(1+H), so the closed loop poles are the solutions s of
1+H = 0:

 
Js 2  K p  Q ' s  K i  0

If the pole-pair has frequency 0 and damping , then


0 
Ki
, 
K p  Q ' 
J 2J0

Hence the PI gains are

K i  J02
K p  2J0  Q '

A damping factor of  = 1/2 is selected for critical damping, and 0 is calculated by setting
the cross-over frequency c to a defined value, e.g. c = 1 rad/s. The cross-over frequency is
the frequency at which |H| = 1, so 0 can then be found by solving

K  i c K p 
 1
i


i c Q '  Ji c

The PI gains are then fully defined.

For the pitch PI controller, the plant can be represented by

Js  Q '   Q ' 

The plant P is now the transfer function from pitch angle to rotor speed, i.e.

 Q '
P
Q '  Js

or  Q  times the torque loop plant. The same analysis can then be carried out, except that
'

now

K i   J 02 / Q '
2J 0  Q '
Kp  
Q '

The pitch PI gains are calculated for each above-rated operating point and a least-squares
procedure is used to calculate a gain schedule to fit for both gains simultaneously, ignoring
points very close to fine pitch where a maximum gain limit is imposed.

Note: the PI gains in Bladed are defined in terms of measured generator speed, and the
torque PI gains are also defined in terms of generator torque referred to the high speed shaft.
Therefore the pitch PI gains derived above must be divided by the gearbox ratio, and the
torque PI gains must be divided by the square of the gearbox ratio.

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, these are very basic algorithms, and it is recognised that in reality there is
great variation in the design of controllers for different turbines.

Bladed therefore offers the possibility of incorporating user-defined controllers in the dynamic
simulations. Through a defined interface, a user’s external control program, written in any
language and compiled either as an executable or (preferably) a DLL, can be used to control
the simulation. Like a real controller, the external controller runs on a discrete timestep.
Using this facility, Bladed simulations can use any real control algorithm, and provide a very
useful means for testing new control algorithms.
For example, the user-defined controller may do any of the following:

 Collective or individual blade pitch angle or pitch rate control during any phase of operation
including power production, stops, starts, idling etc.
 Generator torque control for variable speed or variable slip turbines
 Switch between generators for two-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 an external user-defined controller. In an ideal
situation, the principal algorithm code modules could be shared between the Bladed external
controller code and the code in the actual controller hardware used on the turbine.
Alternatively, it is possible to use a Bladed external controller which communicates directly
with the turbine controller hardware. Bladed also incorporates an option to run in real time,
which allows it to be used as a virtual turbine for testing the real controller.
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.
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.

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. If the built-
in variable speed power production controller is used, the torque controller continues to
1
operate normally until this point .

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, 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

1
Simulations using dtbladed.exe version 2.85 or earlier with the built-in variable speed controller used a
simpler strategy in which the generator was disconnected as soon as the minimum generator speed was
reached.
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.

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

Applied torque*
Tow
Controlled er
torque Yaw spring
Friction Damper
*(Aerodynamic and inertial yaw torque)

Yaw control type Demanded Yaw spring and Controlled


yaw rate damper Friction torque
None No No No No
Rigid Yes No Yes No
Flexible Yes Yes Yes No
Controlled torque No No Yes 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 V 0 and the
instantaneous gas volumes either side of the yaw motor are v1 and v2 then the opposing
torque Qk is given by
  
 V0  V0  
Qk  KP0       
  v1   v2  

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.
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 (Chapter 9).
 Upwind turbine wake: full or partial immersion of the turbine rotor in the wake of another
turbine operating further upwind (Chapter 9).

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 (Chapter 10).
 IEC transients: this option uses wind speed and direction transients as defined by the IEC
1400-1 standard [4.8, 7.7]. It is intended for evaluating specific load cases, for example
during extreme gusts.

7.1 Wind shear

Wind shear is the variation of steady state mean wind speed with height. Three 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:
 log(h / z0 ) 
V (h)  V (h0 ) 
 log(h0 / z 0 
)

7.1.3 User-defined model


A different shear profile may be defined by entering a lookup table giving a wind speed
multiplication factor as a function of height.

7.2 Time varying wind

Various forms of temporal variation of wind speed and direction may be superimposed on the
spatial variations described in Section 7.1 above and Chapter 9.

7.2.1 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.

7.2.2 3D turbulent wind


A 3-dimensional turbulent wind field is generated, with statistical properties representative of
real atmospheric turbulence. Chapter 10 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)
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,).

7.2.3 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 [4.8, 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 7.1.

The actual wind speed at radius r, azimuth  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.

12.5

Half wave
Y0 + 12
A

11.5

11

Full wave
10.5

Y0 10

9.5
IEC edition 2

9
-0.2 t00 0.2 0.4 0.6 0.8
t0 1+ T 1.2

Time

Figure 7.1: Definition of IEC sinusoidal transients


8. MODELLING WAVES AND CURRENTS
For wind turbines sited offshore, the fatigue loads and extreme loads experienced by the 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.

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].

 f 
5
  f   
4

S  f    2 H T   exp 1.25  
2
f 
s p
 f  
 p   p 

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

0.0624
2 
0.185
0.230  0.0336 
1.9  
  f  
2

  1  
 
  exp  0.5
fp  
   
   
   
and
  0.07 for f  fp
  0.09 for f  fp

The Pierson-Moskowitz spectral density function may be regarded as a special case of the
JONSWAP spectrum with   1.0 :

 f 
5
  f  
4

S  f   0.3123H T   exp  1.25  


2
f 
s p
 f  
 p   p 

If the JONSWAP / Pierson-Moskowitz option is selected, the user is required to enter values
for H s , T p 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.
8.3 Wave diffraction approximation

When the wavelength of waves impacting on an offshore structure becomes comparable with
the dimensions of the structure (e.g. the diameter of a monopile), diffraction effects begin to
occur. Waves which have wavelengths much smaller than the diameter of the structural
member do not contribute much to the net force because regions of positive and negative
velocity are experienced by the member at the same time. Bladed provides two different
approaches for including these effects: an approximation based on McCamy-Fuchs theory,
and a simple frequency cut-off.

8.3.1 MacCamy-Fuchs approximation


Explicit models of the diffraction phenomenon are highly specialised and computationally
expensive, and it is therefore not generally feasible to incorporate them in time domain design
tools. An alternative approach is to account for the wave diffraction effect by altering the
hydrodynamic loads experienced by the structure via revised values of the inertia and drag
coefficients Cm and Cd, calculated as a function of the wave frequency. This approach,
described by MacCamy and Fuchs in [8.10], is only applicable to the frequency domain as it is
not possible to alter the Cm and Cd coefficients in time domain analysis. Bladed includes a
variation of this technique in the time domain (MacCamy Fuchs approximation) achieved by
altering the wave energy spectrum rather than the Cm and Cd terms, to give the same resulting
hydrodynamic load on the structure over the full range of incident wave frequencies. In this
procedure, it is assumed that the Cm term dominates the hydrodynamic loading, and so the
wave energy spectrum is altered based solely on how the Cm coefficient varies as a function
of wave frequency.
The procedure followed by Bladed is explained below:
Hydrodynamic force is given by FH = Hydrodynamic Mass Force + Froude-
Krylov Force
= Cm ρ A ẍ + ρAẍ
= ρ A ẍ (Cm + 1)
= ρ A ẍ CM
ρ = water density
A = cross sectional area of cylinder
ẍ = horizontal acceleration
Cm = hydrodynamic mass coefficient
CM = overall hydrodynamic inertia coefficient = Cm + 1

In the case of a vertical cylindrical cylinder with no diffraction effects, Cm = 1. Therefore in the non-
diffraction case, CM = 2.
In the MacCamy Fuchs approximation, the hydrodynamic inertia coefficient CM has the following form:
2
CM = 4A(kro) / π(kro)
Where:
k = wave number
ro = radius of cylinder
2 2 -1/2
A(kr0) = (J1’ (kr0) + Y1’ (kr0))

This reduces CM as k increases. As k tends to zero, CM tends to a value of 2, equal to the non-
diffraction case.

Since it is not possible to modify CM in the time domain within Bladed, CM is held fixed and the
change in hydrodynamic forcing predicted by the MacCamy Fuchs approximation is
reproduced by applying an identical functional form to the entire Hydrodynamic force
equation, considering it to be a modification of ẍ rather than CM.
This modification to ẍ is produced via modifying the wave energy spectrum. The standard wave
2
energy spectrum is the Pierson Moskowitz distribution, in which S(k) α H . In sinusoidal wave theory,
ẍ α H (wave height) giving S(k) α ẍ
2

Thus the wave energy spectrum is multiplied by a correction factor:


S(k) modified = Correction factor × S(k) unmodified
The correction factor is the square of the normalised MacCamy Fuchs C M function. Normalising CM
against the non-diffraction CM value of 2 we have:
2
CM normalised = 4A(kro) /2 π(kro)
2 2
Therefore the correction factor = ( 2A(kro) /π(kro) )
Thus:
2
S(k) modified = (normalised CM value) × S(k) unmodified
2
S(k) modified = (2A(kro) /π(kro) ) × S(k) unmodified

8.3.2 Simple cut-off frequency


Applied forces are calculated from the wave particle kinematics at the member centreline,
using a spectrum in which the high frequencies have been cut off. The frequency cut-off is
based on experimental work by Hogben and Standing [8.2] which shows 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
For a monopile, 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.
Alternatively the cut-off frequency may be specified by the user.

8.4 Wave Particle Kinematics

In the case of Irregular Waves and Regular Waves (Linear Airy option), water particle
kinematics are based on linear Airy theory. The following equations describe the wave particle
 
velocity vector uw  uwx , uwy , uwz , 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  :
H
u wx  cos  w coshk d  z cos  t 
2 sinh kd 
 H
u wy  sin  w coshk d  z cos  t 
2 sinh kd 
H
u wz  sinhk d  z sin  t 
2 sinh kd 
2H
u wx  cos  w coshk d  z sin  t 
2 sinh kd 
2H
u wy  sin  w coshk d  z sin  t 
2 sinh kd 
2H
u wz  sinhk d  z cos  t 
2 sinh kd 
gH
p coshk d  z cos  t 
2 coshkd 

 cos  t 
H
2

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

  kx cos w  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:

 d   t  
z   z    t 
 d 

where  t  is the surface elevation above the location in question.

8.6 Simulation of Irregular Waves

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

 Wave elevation at the support structure centre-line,


 Wave particle velocities, accelerations and dynamic pressures at various points on the
structure,
 Wave forces on the submerged support structure.

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 flow fields, 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 i 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:

S  f m f
z m ,r  Rr  f m  expi r  f m 
4N
z m, r  zm, r

where
f m  m f
f  f max N
and m is in the range 0m N .
The filter weights are then obtained as the transform of the expression:

N
   imn 
wn,r  
m   N 1
zm,r exp
 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
random number is introduced at element 1 and the multiplication and summation process is
repeated.

8.7 Simulation of Regular Waves

Bladed offers two methods for modelling regular waves. Linear Airy waves are as described
in Section 8.4; again Wheeler stretching is used. If the Stream Function option is selected, 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
 ( x, z )  z  n2 X n sinh( nk ( z  d )) cos(nkx)
N

T
where
X 1  wavelength
X n 1   n

and N is the order of the stream function solution.

The stream function  as defined above satisfies the requirements that (i) the shape of the
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 and v
y x

and the dynamic pressure is calculated using Bernoulli’s 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
between the vorticity  and stream function  is the same for the combined flow as for the
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
    f ( )
x 2 y 2

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-Jacotin’s 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
rectangular grid in the x, plane. Treating x and  as the independent parameters, the
velocity components are now given by:

y
u
1
and v x
y y
 

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 Bernoulli’s equation.
Reference [8.5] should be consulted for further details of this method.
8.8 Constrained Waves

Bladed allows the user to include one constrained wave in each irregular wave simulation.
The simulation is otherwise irregular, but the user can constrain it to have a wave of desired
height at a specified time. Bladed uses “NewWaves” to modify the irregular sea around the
specified time so that the constraint is met. A NewWave, by definition, describes the most
likely surface elevation around a peak (or trough). It can be computed from the spectrum:

 ( )   ( )

16
 ( )  2  S ( f ) cos(2f )2df
Hs 0

Here α gives the elevation of the crest, and τ is the time, measured relative to the time of the
crest. The water kinematics below the surface follow the usual linear wave relationships. In
the context of an irregular sea, Bladed uses the NewWave as a tool to achieve a constraint.
A linear combination of a standard irregular sea and a NewWave (centred on the constraint
time) is used to achieve a crest of the desired height at the desired time. The time derivative
of the water surface at the constraint time needs to be set to zero (so that it forms a crest);
this is done by using a slope NewWave:

 ( )   ( )

16
H s 0
 ( )  2
S ( f ) sin(2f )2df

The slope NewWave has zero amplitude but slope α (at the origin), so it can be combined
linearly with the irregular sea to set the time derivative of the surface elevation to zero at the
constraint time.

When the Linear NewWave option is used, Bladed carries out the above procedure and
iterates the height of the crest, so that the difference in elevation between the crest and the
lowest point up to one period either preceding or following the crest meets the user’s
specification.

When the non-linear constrained wave option is used, the stream function solution (see
Section 8.7) is used for one period around the constrained crest, beyond which the linear
irregular model is used. NewWaves are used to constrain the linear sea in such a way as to
allow the two separate models to blend smoothly together. This option combines the
irregularity of the linear model with the full non-linearity achieved by the stream function
model. A full description of this approach can be found in reference [8.11].

8.9 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:


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

8.9.1 Near-Surface Current


The near-surface current velocity profile varies linearly with depth from a specified velocity at
the surface to zero velocity at the reference depth.

8.9.2 Sub-Surface Current


The sub-surface current velocity profile is of the form:

 z  d  
ucs  z     u s 0  z  0
 d  

for 0  z  d , where d is the water depth and u s 0  z  0  is the velocity at the sea surface.
The standard power law exponent  is normally taken as 1/7, but can be changed by the
user.

8.9.3 Near-Shore Current


The near-shore current velocity has a uniform profile, independent of depth.

If desired, a suitable current velocity at the location of the breaking wave may be calculated
as:

ucn  2s 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:

b
HB 
1 a

d B gTB2

where:
a  441  exp 19s 
b  1.6 1  exp 19s 

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 .
8.10 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  u w  uc  u s
u t  u w  u s

8.11 Applied Forces

Having evaluated the total particle kinematics relative to the tower, the resulting forces are
calculated as the sum of two components:

 Dand inertia forcescalculated using the relative motion form of Morison’s equation,
 Longitudinal pressure forces.

These forces are then used to calculate the tower modal forces as described in Section 3.2.2.

8.11.1 Relative Motion Form of Morison’s Equation


To calculate the forces on the tower, the monopile is approximated by 10 cylindrical sub-
elements of equal height. The forces on each sub-element, acting normal to the cylinder
axis,are calculated using the ‘relative motion’ form of Morison’s equation:

 D2  D2
F  Cm  1
1
Lu t   Lu w  Cd  DLut ut
4 4 2

where F is the normal force on a segment of cylinder of length L and diameter D,  is the
water density, C m is the inertia coefficient and C d is the drag coefficient.

8.11.2 Longitudinal Pressure Forces on Cylindrical Elements


Morison’s equation gives the force on a cylindrical element normal to the element’s axis. In
situations where the element is tapered, pressures acting longitudinally on the changing
cross-sectional area may cause a significant axial force to act on the element.

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
sub-element. For a structure with diameter D a at the top of a sub-element and diameter Db
at the bottom, the longitudinal force acting on this portion of the structure is:

F

  Da2  Db2
p

4
No pressure force is included where the end of the tubular member passes through the free
surface or terminates at the sea-bed.
9. TOWER SHADOW AND WAKE EFFECTS

9.1 Tower shadow

Tower shadow defines the distortion of the steady-state mean flow field due to the presence
of the turbine tower or support structure. Three different models are available: a potential flow
model for upwind rotors, an empirical tower wake model for down wind rotors, and a
combined model which is useful if the rotor yaws in and out of the down wind shadow area.

9.1.1 Potential flow model


This model is appropriate for rotors operating up wind of the tower. The longitudinal wind
velocity component up wind 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 in GL coordinates at a distance x in front of
the tower centreline and y to the side of the wind vector passing through the centreline, the
wind speed in the plane perpendicular to the tower axis, V, is given by:

V ( x, y )  V 0  

 y2  x2  2 xy  D  2

2   V0
,

  
 y2  x2 2 y2  x2   2 

provided the point is far from the tower top. For the variation close to the tower top see
Section9.1.4.

9.1.2 Empirical model


For rotors operating down wind 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 x
behind the tower centreline and y to the side of the wind vector passing through the
centreline, the wind speed perpendicular to the tower axis, V, is given by:

V ( x, y )  AV0

where

 y 
A  1   cos2  
 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 9.1.1. 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.

9.1.3 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 average of the velocity predicted by 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.

9.1.4 Application to multi-member towers and end effects


For multi-member towers, the effective wind speed changes due to each member are
calculated individually, using the chosen method, and summed. The magnitude of the total
change is limited to the magnitude of the original velocity. Both the potential flow and Powles
models only modify the components of velocity perpendicular to the member axis.

The effect close to the end of the member varies smoothly for both potential flow and Powles
models. For the potential flow model, the member is assumed to be a line source with ends at
positions p1 and p2. Hence, in a coordinate system with the z axis pointing along the tower (or
member) the resulting velocity is given by:




 z  p2  y 2  x 2  
z  p2 x 2 

 
 y 2  x 2 2 x 2  y 2  ( z  p )2 2     
1 3
y 2  x 2 x 2  y 2  ( z  p2 ) 2 2
 2 



z  p1  y 2  x 2  
z  p1 x 2 
 ,
     
1 3
 y 2  x 2 2 x 2  y 2  ( z  p1 ) 2 2 y 2  x 2 x 2  y 2  ( z  p1 ) 2 2 
 
 2 xy z  p2  xy z  p2  wh
2

V ( x, y , z )  V 0     D  V
 2  
 2 
     
1 3 0 xy
2
 y  x 2
x 2
 y 2
 ( z  p 2 ) 2 2
y 2
 x 2
x 2
 y 2
 ( z  p 2 ) 2 2

 2 xy z  p1  xy z  p1  
  ,
 
 y 2  x2 2 x2  y 2  ( z  p )2 2     
1 3
y 2  x 2 x 2  y 2  ( z  p1 ) 2 2
 1 
0 
 
 
ere V0xy is the magnitude of the velocity in the x-y plane.

For the Powles model, the end effects are taken account of by simplifying it to a diffusion
equation. For line source in the diffusion equation, the fall off at the ends of the source is
proportional to the error function:

  y    2z  p2    2z  p1   


 
A  1   cos2    erf
  WD   erf
 WDT     
 T   WD T 

9.2 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 up wind, 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:

 
2
r


V  V0 1  e 2W 2 
 
 
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 upstream 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.

9.2.1 Eddy viscosity model of the upstream turbine wake

The eddy viscosity wake model is a calculation of the velocity deficit field using a finite-
difference solution of the thin shear layer equation of the Navier Stokes equations in axis-
symmetric 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 9.1.

Incident
flow profile Wake profile

Figure 9.1: 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  (r uv)
U V 
x r r r

The turbulent viscosity concept is used to describe the shear stresses with an eddy viscosity
defined by [7.11]:
 (x)  Lm (x).Um (x)
U
and  uv  
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. The governing
differential equation to be solved becomes:

U U   ( rU / r )
U 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]:

  FK1 Bw (U i  U c )   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]:

 amb  F .K k 2 .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]:

 C t  0.05  (16Ct  0.5)Iamb /1000


Uc
D mi  1 
Ui

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

3.56Ct
Bw 
8Dm (1  0.5Dm )
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  
2

Dm,r 
 exp  3.56 
  Bw  

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

0 .5
W  Bw
3.56

9.2.2 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 I amb, 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  5.7Ct I amb


0.7 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


2 2

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

n r0
xn 
 dr 
 
 dx 

where

m 1
r0  R
2
1
m
1  Ct

n

0.214  0.144m 1  0.134  0.124m 
 
1  0.214  0.144m 0.134  0.124m

and dr/dx is the wake growth rate:

2 2 2
dr  dr   dr   dr 
      
dx  dx   dx  m  dx 

 dr 
   2.5 I 0  0.005 is the growth rate contribution due to ambient turbulence,
 dx 

 dr  1  m  1.49  m is the contribution due to shear-generated turbulence,


  
 dx  m 1  m  9.76
 dr 
and    0.012 B  is the contribution due to mechanical turbulence, where B is the
 dx  
number of blades and  is the tip speed ratio.
10. TURBULENCE MODELS
Bladed incorporates a method for generating simulated three-dimensional turbulent flow
fields which can impinge on the rotor and support structure.

10.1 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 turbulence spectral characteristics, and each pair of time histories has the
correct cross-spectral or coherence characteristics.

Another approach due to Mann [7.18, 7.19] is also included, being now referenced in the third
edition of the IEC standard [7.20].

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 auto-
spectral 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].

It should be remembered, of course, that all these models tend to be based largely on
observations for flat land sites.

10.1.1 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) 4n~u



 u2 (1  70.8n~u2 )5 / 6

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

n x Lu
n~u 
U
x
Here Lu 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) 4n~i (1  755.2n~i2 )



 i2 (1  282.3n~i2 )11/ 6

where

n xLi
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 auto-
spectrum, is:

Cu (r , n)  0.994( A5 / 6 (u )  12 u A1/ 6 (u ))


5/3

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

 nL (r , n) 
2
r
u  0.747 1  70.8 u 
Lu (r , n)  U 

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

( yLu y) 2 ( zLu z ) 2


Lu (r , n)  2MIN (1.0,0.04n 2 / 3 )
y 2  z 2

where y and z are the lateral and vertical components of the separation r, and Lu and Lu
y z

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
2.869 i2  1

4.781 i2 A5 / 6 (i )  A11/ 6 (i ) 
where

 nL (r , n) 
2
r
i  0.747 1  70.8 i 
Li (r , n)  U 

and
i Li (r , n)
i 
r
for i = v or w.

In this case the local length scales are given by:

( yLv y / 2) 2 ( zLv z ) 2
Lv (r , n)  2MIN (1.0,0.05n 2 / 3 )
y 2  z 2
and

1/ 2 ( yLwy) 2 ( z Lw z / 2) 2
Lw (r , n)  2MIN (1.0,0.2n )
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.

10.1.2 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. It embodies a model of the atmospheric boundary layer, and
might therefore be of dubious validity under water. The autospectral density for the
longitudinal component of turbulence is given by:

nSuu (n) 2.987n~u / a 1.294n~u / a


 1  
   
F
 u2 1  2nu / a  1  nu / a 
2 5/ 6 2 2 5/ 6 1
~ ~

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

n x Lu
n~u 
U
x
Here Lu 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) 2.987(1  (8 / 3)(4n~i / a) 2 )(n~i / a) 1.294n~i / a


 1  
   
F2
 i2 1  4n~i / a  1  2n~i / a 
2 11/ 6 2 2 5/ 6

where

n xLi
n~i 
U
and i is either v or w.

The five additional parameters a, 1, 2, F1 and F2 are defined as follows:


F1  1  0.455exp  0.76n~u / a 
0.8


F2  1  2.88 exp  0.218n~i / a 
0.9

 2  1  1
1  2.357a  0.761
a  0.535  2.76(0.138  A) 0.68

where

A  0.115[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:  is the angular speed of rotation of
the earth, and  is the latitude)
u  0.4U  34.5 f .z  / ln( z / z0 )
*

z0  surface roughness length

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
7.5 (0.538  0.09 ln( z / z0 )) p u *
u 
1  0.156 lnu * /  f .z0 

Iu   u / U (the longitudinal turbulence intensity)


  z  
I v  I u 1  0.22 cos4    (the lateral turbulence intensity)
  2h  
  z  
I w  I u 1  0.45 cos4    (the vertical turbulence intensity)
  2h  

although these may be changed by the user for any particular simulation.

The nine turbulence length scales are also defined, as follows:

x
Lu 
 
A1.5  u / u * z
3

2.5K 1z .5 (1  z / h) 2 (1  5.75z / h)
y
Lu  0.5x Lu 1  0.46 exp 35( z / h)1.7 
z
Lu  0.5x Lu 1  0.68 exp 35( z / h)1.7 
x
Lv  0.5x 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

where


K z  0.19  (0.19  K 0 ) exp  Bz / h 
N

K0  0.39 / R 0.11

B  24R 0.155
N  1.24R0.008
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 )  12 u A1/ 6 (u ))


5/3

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

2
 0.747r   2nr 
2

i      c  for i = u
 2 Li   U 

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

( yLu y) 2 ( z Lu z ) 2
Lu (r , n) 
y 2  z 2

while

1.6(r / 2 Lu ) 0.13
c  max(1.0, )
0b
with
b  0.35r / 2 Lu 
0.2

and
2
 0.747r   2nr 
2

0      
 2 Lu   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
2.869 i  1
2

4.781 i2 A5 / 6 (i )  A11/ 6 (i )  for i = v,w

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

i 2 Li (r , n)
i 
r

In this case the local length scales are given by:

( yLv y / 2) 2 ( z Lv z ) 2
Lv (r , n) 
y 2  z 2
and
( yLw y) 2 ( z Lw z / 2) 2
Lw (r , n) 
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.

10.1.3 The Kaimal model


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

nSuu (n) 4n~u



 u2 (1  6.0n~u )5 / 3

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

n L1
n~u 
U
x x
Here L1 = 2.329 Lu where Lu 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

 2 
 n   0.12  
2

C (r , n)  exp  8.8r     
  U   L(r , n)  
 
A three-component Kaimal model has been introduced for compatibility with the IEC standard
1400-1 edition 2 [7.1]. The scale parameter 1 defines the characteristics of the turbulence,
through the following relationships:

Lu = 8.11, Lv = 2.71, Lw = 0.661


x x x

x
n Li
n~i  i = u, v, w
U

nSii (n) 4n~i



σ i2 (1  6.0n~i )5/3

For the longitudinal component,

 2 
 n   0.12  
2

C( r,n)  exp  Hr    
  U   Lc  
 

where the coherence decay constant H = 8.8 and the coherence scale factor Lc = 3.51. The
standard does not define the coherence for the other two components, so the following
expression is used:

 n
C( r,n)  exp  Hr 
 U

A more general formulation for the Kaimal model has also been introduced in which the
parameters Lu, Lv, Lw and Lc can be specified separately instead of specifying 1, and the
x x x

parameter H and can also be specified. This can be used for compatibility with the third
edition of the IEC standard [7.20], which gives H = 12 and Lc = 8.11.

10.1.4 The Mann model


This model is also referred to in the third edition of the IEC standard [7.20]. It is based on a
three-dimensional spectrum tensor representation derived from rapid distortion of isotropic
turbulence by a uniform mean vertical velocity shear. The theory is given by Mann [7.18,
7.19]. The method derives the spectral density for any three-dimensional wavenumber vector,
and all three components of turbulence are then generated simultaneously by summing a set
of such wavenumber vectors, each with the appropriate amplitude and random phase.

This is in many ways quite an elegant approach, but there are some practical limitations to be
aware of. The summation requires a three-dimensional fast Fourier transform (FFT) to
achieve reasonable computation time. The number of points in the longitudinal, lateral and
vertical directions must be a power of two for efficient FFT computation. In the longitudinal
direction, the number of points is determined by the length of time history required and the
maximum frequency requested, and is therefore typically at least 1024. The maximum
wavelength used is the length of the file (mean wind speed multiplied by duration), and the
minimum wavelength is twice the longitudinal spacing of points (which is the mean wind speed
divided by the requested frequency). In the lateral and vertical directions, a much smaller
number of points must be used, perhaps as low as 32, depending on available computer
memory. The maximum wavelength must be significantly greater than the rotor diameter,
since the solution is spatially periodic, with period equal to the maximum wavelength in each
direction. The number of FFT points then determines the minimum wavelength in these
directions. With a realistic number of points, the resulting turbulence spectra are deficient at
the high frequency end. Mann [7.19] suggests that this is realistic, because it represents
averaging of the turbulence over finite volumes of space which is appropriate for practical
engineering applications. Bladed will report the loss of turbulence intensity due to this effect,
so if a certain turbulence intensity is requested for a simulation, the actual turbulence intensity
will be slightly lower due to a loss of high frequency variations.

Note also that the time histories at each grid point may have individual spectra and variances
which can differ from one point to another. This may well be realistic of course, but it means
that if the spectrum or variance is computed at any single point, for example at the hub
position, the result may again be a little different from the expected value.

10.1.5 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, Lfield, 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, Lfield 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). Alternatively there is an option to allow the
turbulent flowfield to wrap around at the end, starting again from the beginning. This allows
an arbitrarily long simulation.
 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. The flow field can also be made tall enough
to encompass the entire rotor and tower.
 The number of grid points in the lateral and vertical directions should be chosen to achieve
adequate sampling of the wind speed variations in the rotor plane. A grid point spacing of
6 – 7 m is likely to be adequate for this. The time taken to generate the turbulence file
increases with the fourth power of the number of points, so it is important not to use more
points than necessary. In the along- wind direction, a resolution of 10 Hz is likely to be
sufficient. For certification calculations, definitive guidelines should be sought from the
certifying body.
 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
which was specified for the simulation since this assumes that the whole time history will be
used. Note also that the number of points along the wind direction 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 along-wind spacing of points will automatically be decreased.
 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 7.2.3 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.
11. MODELLING EARTHQUAKES

For turbines situated in seismically active areas, it is important to assess the loading due to
earthquakes. Bladed offers the user the option of performing calculations using a recorded
time history, or of generating a synthetic accelerogram that satisfies a given Response
Spectrum.

11.1 Calculation of dynamic response

For stiff foundations, Bladed assumes the soil to be rigid and so not to interact with the
dynamic response of the structure.

The equation of motion for a single modal degree of freedom, assuming no coupling with
other degrees of freedom, is as follows: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ii qi  i q  li xg


2

where:

qi is the time dependent modal displacement,


xg is the ground acceleration,

and:
H
li   m(h)i (h)dh is the earthquake participation factor.
0

li xg is added to the generalised forces within Bladed.

With flexible foundations, the seismic acceleration defines the ground motion, which affects
the forces and moments developed in the foundation.

11.2 Response Spectrum

The equation of motion for a single degree of freedom system subject to the  mug force is
as follows:

u  2u   2u  ug

where  is the circular frequency of the structure in rad/s, and  is the damping ratio. This
equation is solved for many single degree of freedom structures having different periods,
each subjected to the same ground acceleration, and each with the same damping ratio. The
plot of maximum response against structural frequency is called the response spectrum [9.1].
Design response spectra are normally presented as a plot of the maximum acceleration
against the period (the inverse of the natural frequency of the structure).
11.3 Generation of response spectrum compatible accelerogram

The algorithm that generates the response spectrum compatible accelerogram is an iterative
procedure. The ability to specify a random number seed ensures repeatability of results. The
process is outlined as follows:

The first step is the generation of a set of random numbers, which are then scaled as a
function of their position in time to create the characteristic earthquake shape. The user is
able to choose between an overall shape typical of stiff or soft soil properties. If soft soil
properties are selected, the stationary part and the decay constant can also be specified. The
time history is then passed through a low and high pass band filter to remove unwanted
frequencies.

After this initial set up has taken place, the process is an iterative one [9.2], with the following
steps:
 Compute response spectrum.
 Compare against Target Response Spectrum at a number of frequency points.
 Check to see if convergence criteria have been met.
 Scale in the frequency domain.

At each iteration, the response spectrum is computed, and compared against the Target
Response Spectrum.

The user is able to specify upper and lower limits of the acceptable deviation of any of these
computed points from the Design Response Spectrum. If all the points are within this range,
the convergence criteria have been met, and the process is complete. If any points are
outside this range, the Fourier transform of the time history is scaled with the following
relationship:
 S aD 
Vg ( new )  Vg ( old )  
 Sa 
D
where Vg is the Fourier transform of the acceleration time history, S a is the target spectrum

value and S a is the actual spectrum value.


The final stage is to correct the peak acceleration, so that it is always equal to the zero period
value on the Target Response Spectrum. The mean is also corrected, so that the velocity of
the ground at the end of the earthquake is always zero.
12. 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 post-
processing calculations is described in this section.

12.1 Basic statistics

The following basic statistical properties of a signal are calculated:

Minimum MIN(x)
Maximum MAX(x)
Mean x
Standard deviation   ( x  x )2
Skewness ( x  x )3 /  3
Kurtosis ( x  x )4 /  4

12.2 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 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.

12.3 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 7.2.3.

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 turbine load 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 12.2). It is generally a good approximation to assume that the stochastic part of the
load is Gaussian, so its probability distribution is:

1 / 2 x2
ex
2
p( x) 
 x 2

where x is the standard deviation of x. For such a signal, Rice [10.1] has derived the
probability distribution of signal peaks as:

  
  
1    2 / 2 (1 2 )   2 / 2 
2
  
pˆ ( x)  e  e  1  erf  
 x 2 2 x   22  2  
   
  
where

  x / x
   0 / m
M2
0  (the zero up-crossing or apparent frequency)
M0
M4
m  (the frequency of peaks)
M2

M i   f i H ( f )df (the ith spectral moment)
0
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

pˆˆ ( )  Npˆ ( )1  Q 


N 1

where

Q( )   pˆ ( )d , and

N  Number of peaks in the period.


Davenport [10.2] combined this with Rice’s equation to give the following analytical
expression for the probability distribution of extremes:

pˆˆ ( )   e 
where
   0T e
2
/2
and
T  time period.
The mean of this distribution is


ext   

where
  2 ln( 0T ) and
  0.5772 (Euler’s constant).

As the term  0T increases, the distribution of extremes has a larger mean and becomes very
narrow.

For an operational turbine whose loads are a combination of stochastic and periodic
components, Madsen et al [10.3] proposed an approach based on Davenport’s 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:
z
t1
zmax

zmean t2

Time

t3
zmin

T0

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

 
ye max  zmax   x  1  
 1 

 e max   x
6 1
where
1  2 ln(1 0T )
t1  z2
1  
T0 ( zmax  zmean )( zmax  zmin )

while for extreme minima:

  
ye min  zmin   x   3  
 3 

 e min   x
63
where
3  2 ln( 3 0T )
t3  z2
3  
T0 ( zmean  zmin )( zmax  zmin )

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.

12.4 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: (1  cos(2f )) / 2
(d) Hamming: 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: If checked, a linear trend is calculated for each segment and removed from it
before windowing, this is usually desirable. If left unchecked, the mean is calculated instead
for each segment and removed from it before windowing.

12.5 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 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.

12.6 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.

12.6.1 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 is as defined in Recommended Practices for Wind Turbine
Testing and Evaluation, 3. Fatigue Loads, IEA, 2nd edition 1990. It 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 peak-to-peak
amplitude (i.e. the range) of a 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 S i m  m
 i 
 
 Tf 
 
where ni is the number of cycles in load range Si and T is the duration of the original time
history.

12.6.2 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 [10.4] linear cumulative damage law. Failure will occur when
the “accumulated fatigue damage number” is equal to 1.0 as follows::

ni
Ni
 1.0
i
th
where ni is the number of rainflow cycles of the i 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:

1 1
log S  log k  log N
m m
so that:

N  kS  m
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
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.

12.7 Annual energy yield

The annual energy yield is calculated by integrating the power curve for the turbine together
with a Weibull or user-defined 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:


k
 V 
 
F (V )  1  e  cV 

where F is the cumulative distribution of wind speed V. Thus the probability density f(V) is
given by
k
 V 
V k 1  cV 
f (V )  k e
(cV ) k
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:

 1
c  1 / 1  
 k

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.

Frequently a steady state power curve is used, combined with a 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 distribution to use
in this case would be one representing the distribution of 10-minute mean wind speeds in a
year. For a Weibull distribution, this will typically have a slightly smaller shape factor than that
for hourly means.

12.8 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.

12.9 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 [10.5]. 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 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:
4 2 2 2 2
U + U (2{QX - PR} - U0 ) + (QX - PR) + (PX + QR) = 0

where U0 is the voltage at the infinite busbar, and P and Q are the active and reactive power
respectively.

12.10 Extreme load extrapolation

The Extreme load extrapolation calculation uses statistical methods to predict the extreme
load for a defined return period by extrapolating from the extreme load values obtained from a
large number of simulations performed with a range of external conditions.

Probability distributions are fitted to these extreme load results. These fitted distributions are
then weighted and summed to give an overall response function for each load. Extrapolated
extreme loads are then obtained for the defined return period from these overall response
functions.

Two probability density functions are used for fitting to the extreme data. These are the (2-
parameter) Gumbel distribution and the 3-parameter Weibull distribution. It has been found
[10.6] that these are likely to be the most applicable distributions to wind turbine loading. The
equations below are the cumulative distribution function form of these distributions:

Gumbel:

 x u 
 

F x   e e  s 

3-parameter Weibull:
k
 x u 
 
F x   1  e  s 

where

u = location parameter
s = scale parameter
k = skewness parameter

The method of moments equates sample moments to parameter estimates. The method of
moments has the advantage of simplicity. The disadvantage of the method of moments is that
it tends not to be as accurate in its parameter estimates as the maximum likelihood and least
squares estimators. The moment estimates are used as starting values for the more precise
maximum likelihood and least squares estimates. They are also output as a check.

Maximum likelihood estimation begins with the mathematical expression known as a


likelihood function of the sample data. Loosely speaking, the likelihood of a set of data is the
probability of obtaining that particular set of data given the chosen probability model. This
expression contains the unknown parameters. Those values of the parameter that maximize
the sample likelihood are known as the maximum likelihood estimates.

For the Gumbel distribution, the method of least squares fits a straight line in the usual way
but in Gumbel axes and for the three parameter Weibull distribution, a similar technique is
employed but also minimising the sum of the squares for the third parameter.

There is no way of deciding quantitatively which distribution or fitting method best suits the
data. This motivates the diagnostic plots which can help qualitatively. For more detailed
theory points including an example of when distribution choice is important, please refer to
the GH study into extrapolation carried out under the European Union funded project,
‘Recommendations for the Design of Offshore Wind Turbines’ (RECOFF) [10.7]. Generally the
3-parameter Weibull distribution is more flexible for fitting that the Gumbel distribution due to
the extra skewness parameter. The Gumbel distribution is always useful as a check.

Two updates have been made for version 4.2 and later, as follows:

Update 1 – A change to the short-term probability distribution when using the Peak Over
Threshold (POT) method.
Previously the probability assigned to each local extreme(maxima) in a given simulation was
simply assumed to be equal to the probability assigned to the global extreme(maxima) of the
simulation. Now the probability assigned to each local extreme(maxima) in a given simulation
is the probability distribution of the local extreme(response maxima) raised to the power of the
average number of local extremes(responses) taken from all simulations within the individual
wind bin. This results in a change to the extrapolated loads when using the ‘POT’ method.

Update 2 – A change to the calculation of the exceedance probability of the extrapolated


result.
Previously the exceedance probability is assigned to the period of a single simulation divided
by the normalised period summed by all simulations. Now the exceedance probability is
assigned to the period of a single simulation divided by the period summed by all simulations.
This results in the exceedance probability being less than the one calculated from previous
versions, but this change will have no influence on the extrapolated value.

12.11 Offshore Code Checking

The aim of the ‘Offshore Code Checking’ post processing tool in Bladed is to unite the main
competencies and advantages of Bladed with those of the offshore analysis packages ASAS
or SACS. On the one hand, Bladed is able to perform rigorous non-linear, time domain
simulation of the integrated WTG and support structure model. On the other hand, ASAS and
SACS are customised to perform robust and automatic fatigue and ultimate limit state (code-
checking) routines.

12.11.1 Bladed-ASAS link


The Bladed-ASAS option in the Offshore Code Checking module transfers the description of
the space-frame structure (including its geometry, mass, stiffness, material properties and
hydrodynamic coefficients) and the internal load time-series data from all relevant load cases
simulated into an ASAS-format database. The format of this database is recognised by the
ASAS post-processing tools FATJACK and BEAMCHECK and can thus be accessed and
processed by these tools in their normal manner.

The Bladed-ASAS option also automatically creates a FATJACK or BEAMCHECK input file,
based on a combination of user-defined inputs (pre-defined by the user in a template file
within the Bladed installation directory) and automatic input commands from Bladed. The
latter include transfer of the following parameters from Bladed into the ASAS input file:
- Project name
- ASAS ‘structure’ name
- load case list
- associated probability of each load case
- design lifetime

A detailed explanation of the ASAS input commands is given in the FATJACK and
BEAMCHECK user manuals.

12.11.2 Bladed-SACS link


The Bladed-SACS option in the Offshore Code Checking module transfers the description of
the space-frame structure, load case list, associated probability of each load case, and design
lifetime from Bladed directly to SACS. The SACS code then performs a conversion routine on
the Bladed internal load time-series data to the required format for use by the SACS code-
checking routines.
13. THE WINDFARMER LINK MODULE
13.1 Purpose and motivation

The Windfarmer Link is a software module dedicated to the calculation of loading on wind
turbines operating in specific environmental conditions. The tool unites the respective
capabilities of Windfarmer [11.1, 11.2] and Bladed software to provide an integrated method
of site-specific fatigue load assessment. This represents a major advance, placing the
calculation of site specific wind turbine loads on the same computational basis as the
evaluation of wind farm energy production which is routinely performed for project developers
and investors.

The motivation to determine site-specific fatigue loads follows from the inability of class
structures used in wind turbine certification standards to account for the complexity inherent in
an actual wind farm. Site-specific calculations are therefore required to determine whether the
loading on a turbine operating in a particular environment falls within the machine’s design
envelope. The unique specification of a given site (in terms of its turbulence levels, air
density variation, up-flow angles, wake-flow characteristics, wind probability distribution, and
other parameters) will determine the loading experienced by each turbine and should be
accounted for in any valid model.

Successful selection of turbine and turbine layout for a given site is therefore dependent on
an accurate understanding of the associated site-specific loads, emphasising the need for
effective calculation methods.

13.2 Overview of site-specific fatigue load calculations

The principal aim of a site specific load calculation is to assess which turbines on site, if any,
exceed their certification/design levels. A wide range of load components are investigated.
This is achieved by running a number of time-domain simulations of turbine operation (and in
some cases turbine idling scenarios). It is standard practice to base these simulations upon
the associated fatigue load cases used for the turbine certification load calculations, but
significantly the generic external conditions of the certification standard are substituted with
the specific site environmental conditions. A salient consideration in this regard is the effect of
wake interaction. In summary, the site-specific conditions relevant to fatigue load calculations
include:

 air density
 wind speed and direction probability distribution (wind rose information)
 ambient turbulent intensity distribution
 wind shear
 terrain up-flow angle
 wake interaction

An initial evaluation of the site prior to the simulation stage can be used to select turbines
which are obviously the most severely loaded on site. Candidate turbines can be identified on
the basis of wake interaction (itself determined by the number of neighbouring machines, the
spacing between turbines and the predominant wind direction) and the variation in wind
shear, up-flow angle and annual mean wind speed across the site. In cases where candidate
turbines are not so obviously identified, a larger number of machines should be selected, the
worst-case scenario being an explicit analysis of the whole site.

Section 13.9 looks at the issue of turbine selection in more detail. An approximate pre-
simulation ranking method is described whereby turbines on site are ordered based on the
environmental severity of their location.
Once the study turbines have been selected, the WindFarmer Link module is used to
construct Bladed project files for creating the wind, performing time domain simulations, and
post-processing the results. A Bladed batch can then be constructed which incorporates all
these calculations.

In common with certification load calculations it is important to run time-domain simulations to


model the influence of non-linearities introduced by the rotor aerodynamics and the turbine’s
control systems.

The final stage is to compare the calculated site-specific fatigue loads (normally expressed as
a lifetime damage equivalent load) with the design constraints of the turbine’s components,
including safety factors. If site specific loads are found to exceed design loads for any given
turbines, repositioning of turbine locations and/or shut-down procedures may be considered
and the loads re-calculated for the revised specification.

13.3 General methodology of the Windfarmer Link

Windfarmer is used to provide a detailed specification of a given site in terms of its ambient
and wake-affected turbulent intensity levels, air density, up-flow angles, wind speed and
direction distribution and wake-flow characteristics.

The Windfarmer Link module is used to manipulate the site information from Windfarmer
(under the guidance of user-defined constraints) to generate wind file specifications, multiple
time-domain simulations and post-processing calculations in Bladed that then represent the
particular operating environment of selected turbines.

Bladed is used to create the wind-files, perform the simulations and carry out post processing
in order to generate results in terms of lifetime equivalent loads and/or rainflow cycle counts.

13.4 Wake modelling

It is well known that the fatigue loading on a wind turbine operating in a wind farm is critically
affected by the downstream wake flow of neighbouring machines. Two load-increasing
mechanisms are observed:

 an increase in the turbulent intensity of the incident flow


 the mean profile of velocity deficit in the wake through which the blades of the down-
stream turbine pass

The nature of both mechanisms is highly dependent on the separation between incident and
upstream turbine.

Two widely used methods to account for wake effects on fatigue loading are the Frandsen
formulation [11.3] and the ‘Dynamic Loads in Wind Farms II’ (DLWF) method [11.4]. Both
introduce the concept of an artificial design turbulence to represent the complex mechanisms
of wake-affected fatigue loading.

Departing from the concept of design turbulence, the WindFarmer Link method draws on the
capability of a validated wake flow modelling tool conventionally used for wind farm energy
calculations and couples it with the customised load calculations of Bladed to provide a state-
of-the-art tool for site-specific load assessment. An important feature of this method is its
ability to model the mechanisms of increased turbulent intensity and mean velocity profile
explicitly.

13.4.1 Wake affected turbulence intensity


In contrast to the concept of design turbulence invoked in the Frandsen and ‘Dynamic Loads
in Wind Farms II’ methods, the turbulence intensity values calculated by WindFarmer and
subsequently used in the simulations of the WindFarmer Link method are what one should
expect to measure physically on site.

In WindFarmer, wind farm turbulence levels are 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, the turbine
thrust coefficient, the distance downstream form the rotor plane and the length of the near
wake:

I add  4.8Ct I amb


0.7 0.68
x / xn 0.57
Where xn is the calculated length of the near wake using the method proposed in [7.16, 7.17].
The characterisation was subsequently amended slightly to improve the prediction, as shown
below [7.15]:

I add  5.7Ct I amb


0.7 0.68
x / xn 0.96
Using the value of added turbulence and the incident ambient turbulence, the turbulence
intensity at any position in the wake can be calculated. The model also accounts for the
turbine not being completely in the wake. The same model is used in Bladed for the Upwind
Turbine Wake model (Section 9.2).

The ambient turbulence intensity is best derived from measurements. Alternatively


WindFarmer can predict the turbulence intensity from an input surface roughness length,
which is representative of the site, using [7.10]:

1
I amb 
lnh / z0 

The turbulence intensity is defined here as the quotient of standard deviation and mean wind
speed at high wind speeds.

In the WindFarmer model set-up the ambient turbulence distribution should be the mean
level. However, for load calculations, it is standard practice to use the characteristic or
representative turbulence. The characteristic turbulence cannot be used directly in
WindFarmer as it would introduce non-physical turbulence values into the wake model. This
problem is negotiated by defining an array of characteristic turbulence factors in the
WindFarmer Link interface which scales on the mean ambient turbulence levels subsequent
to execution of the eddy viscosity wake model. Further details of this functionality are given in
the Bladed User Manual.

13.4.2 Eddy viscosity wake model


There are three wake models available within WindFarmer;

 PARK model,
 Modified PARK model,
 Eddy viscosity model.

When using the WindFarmer Link module, the eddy viscosity model should be used.

Due to the complexity of the wake directly behind the rotor, the eddy viscosity model is
initiated from two diameters downstream. This is assumed to be the distance where pressure
gradients no longer dominate the flow. If a turbine is within this limit, the program resets the
axial distance offset to a value of two diameters. A list of the input parameters to the eddy
viscosity model in WindFarmer can be found in [11.1].

Using this wake model, WindFarmer calculates a Gaussian wake profile and outputs the
centreline velocity deficit, wake width (Bw in Section 9.2.1) and centreline horizontal offset.
The WindFarmer Link module uses this information to calculate the corresponding
parameters required by Bladed to define the Gaussian Wake model for the Upwind turbine
wake.

13.5 Wind flow modelling

WindFarmer uses a wind flow model to ascertain the variation in wind speeds incident on
each turbine at their respective locations on the site. It is conventional to use the WAsP code
for the site wind flow analysis but any wind flow model that outputs the same standard file
formats as defined by WAsP may be used. The model uses Weibull A and k parameters to
represent the directional wind speed probability distribution.

WindFarmer requires the following inputs to perform a wind flow calculation:

1. A WAsP wind resource grid (WRG) file at the turbine hub height with extents covering
all intended turbine locations, or a WAsP RSF file at the turbine locations
2. Turbine locations as grid co-ordinates,
3. Turbine performance data, including power, thrust and rotor speed characteristics.

13.5.1 Site mast data and speed-up factors


WindFarmer allows the user to use the probability distribution of wind speeds measured at
an on-site mast and scale them for each direction using predictions from the wind flow model.

The flow and performance matrices output from WindFarmer are referenced to the
associated site mast wind speed. Variation in wind speed between the site mast and turbine
positions due to topographical effects are then described by the software in terms of speed-up
factors. The speed-up factors vary as a function of direction.

It is therefore a requirement of the WindFarmer Link module to re-scale the flow and
performance matrices so that the dependent variables are given in terms of the turbines’
localised wind speeds. These re-scaled matrices are subsequently used in the generation of
Bladed simulations.

13.5.2 Wind speed and direction distribution


An important aspect of site specific fatigue load assessment is the wind speed and direction
distribution (wind rose) observed at each turbine position. WindFarmer gives the distribution
relative to the associated mast wind speeds. It also internally re-scales the distribution with
the turbine speed-up factors to provide localised wind rose data at each turbine location. The
WindFarmer Link code does not therefore need to re-scale the wind speed and direction
distributions – the user is merely required to select the localised distribution when outputting
the flow and performance matrices from Windfarmer. Note that these local distributions do
not account for wake effects.

The wind speed and direction distribution is important in load calculations for two reasons:

1. The wind direction information determines the amount of time spent operating in a
wake.

2. The wind speed distribution determines the operational hours per year (time between
cut-in and cut-out wind speeds) and the annual mean wind speed at the turbine.
13.5.3 Upflow angles
The localised ground slope at each turbine position on site is calculated in WindFarmer
through the following routine. A circle with radius 2D (where D is the turbine rotor diameter) is
centred on the turbine’s tower base coordinates. At each degree of the circle, the elevation of
the ground directly above or below the associated circumferential point is used in conjunction
with the elevation of the turbine base to calculate the gradient between the two points. The
ground slope is assumed to be constant between these two points. A terrain angle is thus
calculated for each of direction bin around each turbine on site. This is assumed to be equal
to an effective upflow angle.

For fatigue load assessment the WindFarmer Link module only selects simulations at certain
direction sectors. The code effectively combines information for multiple direction bins into a
bin for which a single simulation is defined. Section 13.6 describes this process in detail. With
particular regard to upflow angles, the user has two options:

1. Assign the maximum up-flow angle observed over the combined direction bins to
the simulation.
2. Calculate a weighted mean up-flow angle as follows:

iN iN
 weighted mean   Pi  i  /  Pi
i 1 i 1

where Pi is the probability of the ith direction bin assigned to the simulation,  i is the upflow
angle of the ith direction bin assigned to the simulation and N is the total number of direction
bins assigned to the simulation.

13.6 Selection of wake-affected and ambient simulations

The WindFarmer Link module receives information from WindFarmer in the form of flow and
performance matrices, with mast wind speeds and wind directions introduced as the
independent variables. A number of these matrices, such as those defining the speed-up
factors or upflow angles on site, are one-dimensional as these site parameters do not vary
with wind speed, only wind direction. The majority of matrices output (those defining
turbulence levels, wake profile parameters, wind speed / direction probability distribution and
so on) are two-dimensional, varying as functions of both wind speed and direction. The
WindFarmer Link module can accept any resolution in the number of wind speed bins and
direction bins defined in the WindFarmer simulation. However, the user should be aware that
if a smaller number of bins is selected, a less accurate simulation will result. The wind speed
bins are recommended to go up to 70m/s to ensure that the model is valid over a conceivable
range of speed up factors. Simulations run with less than 72 direction steps should be
avoided.

13.6.1 Wake-affected simulations


For each turbine selected for analysis the WindFarmer Link module scans through the wind
speed and wind direction bins of the turbine’s flow and performance matrices (subsequent to
their speed-up re-scaling) and determines for each bin whether the turbine is operating in
wake-affected or ambient conditions. Any particular turbine is deemed to be wake-affected in
a particular wind condition if any wake offset (the distance between the hub and the wake
centreline) is less than the wake half-width at that point (Bw/2) plus one turbine radius (R): see
Figure 13.1.
R

Wake
offset
Bw/2

Wind direction

Figure 13.1: Wind direction at which a turbine just becomes wake-affected

The user can opt to have either one or two simulations in each wake. If one simulation is
chosen, the wind direction will be chosen so that the wake offset is zero; if two simulations are
chosen, they will be done with wind directions giving wake offsets of +R (one turbine radius).

In real wind farms, the situation will be complicated by the occurrence of overlapping wakes
from different neighbouring turbines. This will be characterised by a range of wind directions
within which there are two or more distinct wakes affecting the turbine without any ambient
direction bins separating them. Then for any particular wind direction Windfarmer will select
the wake characteristics based on whichever wake has the greater velocity deficit at the
turbine position. If the user opts to have one simulation per wake, these simulations will be
defined for each wind direction such that turbine is on the wake centreline of each wake; with
two simulations per wake, wind directions giving wake offsets of +R for each wake are
chosen. Then if any of these simulations is for a wake which is not the strongest wake for
that wind direction, that simulation is discarded.

13.6.2 Ambient simulations


Ambient conditions are defined implicitly as those directions where there are no incident
wakes. The user is asked to constrain the number of ambient simulations per wind speed bin
required in the analysis. Clearly, a smaller number will reduce computation time, but a larger
number may be recommended if wide variation in non wake-related parameters over the wind
direction range (such as ambient turbulence or up-flow angle) is observed on site. Two
methods of selecting the ambient sectors are available.

13.6.2.1 Uniform sectors


The 360-degree direction range is divided into X sectors, where X is the number of
ambient simulations per wind speed bin assigned by the user. The code then determines:

 the maximum ambient turbulence observed in each sector.


 the maximum or weighted-mean up-flow observed in each sector (see section
13.5.3).
 the amount of time spent in ambient conditions in each sector.

If in a given sector no time is spent in ambient conditions (that is, the sector is completely
dominated by wake-affected operation) then no ambient simulation is defined. Hence, for
example, a user may specify 5 ambient simulations per wind speed bin and yet only
observe 3 simulations allocated by the code; the other 2 sectors are completely
immersed in wake-affected flow.

13.6.2.2 User-defined sectors


Alternatively, the user may manually assign ambient simulations. The directional range is
partitioned into twelve 30 degree sectors. The user, via the software interface, can then
assign any permutation of sectors to a variety of simulations.
st
For example, the user may observe that ambient conditions in the 1 sector are similar to
th th th
the 7 , 8 and 11 sectors. Simulation 1 should be assigned to these 4 sectors.
nd rd th th th th th
Furthermore, the 2 , 3 , 4 , 5 , 9 10 and 12 sectors may have similar conditions.
th th
These should be lumped as Simulation 2. Finally, the 6 and 11 sectors, if similar, can
be described by Simulation 3.

Any combination may be implemented. The most simplistic will be assigning all sectors
to Simulation 1. The most detailed will have 12 separate simulations; one for each
sector.

13.6.3 Non-operational simulations


Ambient and wake-affected simulations are only assigned by the code between the turbines’
range of cut-in and cut-out wind speeds. The sum of the wind speed / direction probability
distribution within this operational range will clearly be less than one. The Bladed post-
processing facility for fatigue loads requires that the total amount of time in one year (8766
hours) is assigned to simulations, including time below cut-in and above cut-out. This is
achieved by generating a non-operational idling simulation which effectively contributes no
damage by using values of wind speed and air density very close to zero. The Windfarmer
Link creates this non-operational simulation automatically for each turbine and subsequently
introduces it into the post-processing routine to account for time spent outside the turbine’s
operational range.

13.7 Generation of Bladed simulations

After determining the position of wake-affected and ambient simulations, the code proceeds
to write the associated Bladed project files.

A template file for the simulations is provided by the user through the interface. This file
should contain all the normal information required to successfully run a turbulent wind, time-
domain calculation in Bladed. Furthermore, it is a requirement that the ‘up-wind turbine wake’
module and the ‘idling’ module are both specified in the template. However, the actual
parametric values assigned to these specific modules are, at this stage, unimportant.

The template file as provides the description of the actual turbine structure (rotor, nacelle and
tower assembly), the control system and the dynamic simulation controls, whereas the
processed information from Windfarmer specifies the environmental conditions. Clearly, it is
usually important to ensure that common information assumed in the two models is consistent
– for example, the turbine hub height and rotor diameter.
User-defined Bladed Environmental conditions
simulation template from Windfarmer

 Structural description  Site turbulence


 Modal behaviour  Wake characteristics
 Output options  Air density
 Simulation parameters  Terrain up-flow
 Control system  Wind shear?

Multiple Bladed
site-specific simulation files

Figure 13.2: Schematic showing generation of site-specific Bladed


simulations

The Bladed simulation template may contain values for the environmental parameters; but
these will merely be overwritten during execution of the Windfarmer Link software.

In situations where the user wishes to overwrite particular outputs from Windfarmer, an
options module in the user interface allows particular values of air density, wind shear and up-
flow angle to be assigned to all simulations. See the User Manual for further details.

13.8 Wind file structure


Three dimensional, three component wind turbulence files are created with the “Improved Von
Karman” model (Section 10.1.2).

The wind speed bins specified by the user in the Windfarmer Link screen are used to define
the individual wind speeds and total number of the generated wind files. The turbulence data
from Windfarmer is used to define the turbulent characteristics of the wind files.

The Windfarmer Link module determines the required dimensions of the wind files from the
hub height and rotor diameter information contained in the turbine summary data output from
Windfarmer. It also assigns random number seeds to each wind file.

The turbulence data from the wind file is then combined with the appropriate wake profile,
wind shear and upflow angle during the execution of each dynamic simulation.

13.9 Pre-processing ranking routine

13.9.1 Turbine rank methodology


The highly detailed site-specification offered by the Windfarmer Link method can result in a
large number of required Bladed simulations. The user can vary the total number through
selection of the number of wind speed bins, simulations across a wake sector and ambient
simulations. However, even with these controls, for a large wind farm the sheer number of
simulations required to explicitly analyse the loading on each machine may become
prohibitive, both in terms of computational speed and disk space.

Traditionally engineers performing site-specific assessments have used their understanding


of the site conditions to make an educated assumption about which turbine positions are most
severely loaded and which are relatively benign. This exercise, though beneficial in reducing
processing time, is potentially quite subjective.

The Windfarmer Link module’s pre-processing ranking routine is introduced as a more


quantitative method for determining the most severely fatigue-loaded turbines on site.

The algorithm computes an ‘indication of severity’ parameter X for each turbine T, for a range
of SN-slope values, such that:

1
 i, j m
X T   Pi , j ,T  U i ,T  TI i , j ,T  
m

 

where U is the hub height rotor-averaged wind speed (including wake deficit effects), TI is the
turbulence intensity distribution, P is the wind speed / direction probability distribution, m is
the Wohler exponent or inverse S-N slope (Section 12.6.2), T is the turbine index, i is the wind
speed index and j is the wind direction index of the flow and performance matrices.

The calculation integrates the standard deviation of the wind speed at each turbine with the
probability distribution, with an adjustment for the material inverse S-N slope m. The
probability distribution is not normalised, so that it accounts for the variation in operational
hours across the turbine positions.

The ‘indication of severity’ parameter X should in no way be interpreted as an absolute


measure of load levels; it is merely an indication of how loads may vary relatively between
turbines on site.

The ranking method does not take into account the wind shear and upflow angles, so in
addition to the turbines suggested by the ranking, the user should manually select any
turbines experiencing high wind shear and/or upflow angles for explicit simulation.

13.9.2 IEC class comparison


The user can introduce the generic wind class that the turbine on site has been certified to as
a datum for the ranking calculation. The IEC class is defined through the interface and the
‘indication of severity’ parameter X is calculated for this wind class. The margins between the
site specific rank numbers for each turbine on site and this IEC wind class value are
computed and presented in plots. Note that by convention, a negative margin indicates that
the site specific rank number exceeds the IEC wind class value, whereas if no IEC class
datum is selected, the absolute site specific rank numbers are shown in the resulting plots.

13.10 Post-processing routine

The Windfarmer Link method automatically generates rainflow cycle count simulations for
execution in Bladed. The user is required to define a rainflow simulation template which the
link then alters to correctly assign load cases and their associated probabilities based on the
site specification provided by WindFarmer. Final loads are then extracted from Bladed in the
normal way.
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”, Risø 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

2.11 Kirk Gee Pierce “Wind Turbine Load Prediction using the Beddoes-Leishman Model for
Unsteady Aerodynamics and Dynamic Stall”, University of Utah Thesis, August 1996

3.1 Nikravesh, P.E., Computer-Aided Analysis of Mechanical Systems, Second Edition.


Printice-Hall, Englewood Cliffs, New Jersey, 1988

3.2 Shabana, A.A., Dynamics of Multibody Systems, Second Edition. Cambridge University
Press, Cambridge, 1998

3.3 Géradin, M. and Cardona, A., Flexible Multibody Dynamics. John Wiley & Sons Ltd,
Chichester, 2001

3.4 Meijaard, J.P., A Proposal for the Application of Multibody Dynamics to the Simulation
of Wind Turbines. Technical report No. 2005-1-B prepared in commission of Garrad
Hassan and Partners Ltd, Bristol, December 2005 (Internal document, not publicly
available)

3.5 Cook, R.D., Malkus, D.S., and Plesha, M.E., Concepts and Application of Finite
Element Analysis, Third Edition. John Wiley & Sons, New York, 1989
3.6 Clough, R.W. and Penzien, J., Dynamics of Structures, Second Edition. McGraw-Hill,
Inc., New York, 1993

3.7 Przemieniecki, J.S., Theory of Matrix Structural Analysis. McGraw-Hill, Inc., New York,
1968

3.8 Craig Jr., R.R., ‘Coupling of Substructures for Dynamic Analysis: An Overview’,
AIAA Paper, No. 2000-1573, AIAA Dynamics Specialists Conference, Atlanta, Georgia,
April 5, 2000

3.9 Craig Jr., R.R. and Bampton, M.C.C., ‘Coupling of Substructures for Dynamic Analysis’,
AIAA Journal, Vol. 6, No. 7, pp. 1313-1319, 1968
th
3.10 Kreyszig, E., Advanced Engineering Mathematics, 9 edition, John Wiley and Sons,
inc, 2006

3.11 Bir, G. “Multiblade coordinate transformation and its application to wind turbine
analysis”, ASME wind energy symposium, Reno, Nevada, 7-10 January 2008.

3.12 Hansen M.H., “Improved Modal Dynamics of Wind Turbines to Avoid Stall-induced
Vibrations”, Wind Energ. 2003; 6:179–195.

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.

4.5 Ekanayake J B, Holdsworth L, Wu X, Jenkins N: “Dynamic modelling of doubly fed


induction generator wind turbines”, IEEE Transaction on power system, Volume 18, No.
2, May 2003, pp 803-809.

4.6 Vas P: “Sensorless vector and direct torque control”, Oxford University Press, 1998,
ISBN 19 8564651.

4.7 Krause P C, Wasynczuk O, Sudhoff S D, “Analysis of Electric Machinery and Drive


System”, 2nd Edition, John Wiley press, 2002, ISBN: 0-471-14326-X.

4.8 Kerkman R J, Krause P C, Lipo T A, “Simulation of a synchronous machine with an


open phase”, An International Quarterly on Electrical Machines and Electromechanics,
1977, pp 245-254

4.9 Krause P C, Wasynczuk O, Sudhoff S D, “Analysis of Electric Machinery”, IEEE, 1995,


ISBN: 978-0780311015.

4.10 Duane C H, “Brushless permanent magnet motor design”, 2nd Edition , Magna Physics
Publication , 2006, ISBN: 1881855155.

7.1 IEC 1400-1, Wind turbine generator systems - Part 1: Safety requirements, First
edition, 1994.
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, 1999.

7.8 Ainslie J F, “Development of an eddy viscosity model for wind turbine wakes”,
th
Proceedings of 7 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.

7.18 J. Mann, “The spatial structure of neutral atmospheric surface-layer turbulence”, J. of


Fluid Mech., v. 273, 1994, pp. 141-168.

7.19 J. Mann, “Wind field simulation,” Prob. Engng. Mech., v. 13, n. 4, 1998, pp. 269-282.

7.20 IEC 1400-1, Wind turbine generator systems - Part 1: Safety requirements, Third
edition, 2004.

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 Non-


Uniform 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.

8.10 MacCamy, R.C. and Fuchs, R.A., Wave Forces on Piles: a Diffraction Theory, Tech.
Memo No. 69, US Army Corps of Engineers, Beach Erosion Board, 1954.

8.11 Rainey P J and Camp T R, “Constrained non-linear waves for offshore wind turbine
design”, Journal of Physics: Conference Series 75 (2007) 012067, “The Science of
Making Torque from Wind”,
http://www.iop.org/EJ/article/1742-6596/75/1/012067/jpconf7_75_012067.pdf.

9.1 Gupta A K, “Response Spectrum Method”, Blackwell Scientific Publications, 1990.

9.2 Clough R W, and Penzien J, “Dynamics of Structures”, McGraw Hill, 1993.

10.1 Rice S O, “Mathematical analysis of random noise”, Selected papers on noise and
stochastic processes, ed. N Wax, 1959.

10.2 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.

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


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

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


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

10.5 Flickermeter functional and design specification, BSEN60868, 1993, and evaluation of
flicker severity, BSEN60868-0, 1993, equivalent to IEC 868-0, 1991.

10.6 Cheng, P.W., “A Reliability Based Design Methodology for Extreme Responses of
Offshore Wind Turbines”, PhD thesis, Delft University of Technology, 2002.

10.7 Norton, E. J. “Investigation into IEC Offshore draft standard Design Load Case (DLC)
1.1”, GH report 2762/BR/20, or
http://www.risoe.dk/vea/recoff/Documents/Sec_3/RECOFFdoc059.pdf
11.1 “WindFarmer Theory Manual”, GH & Partners Ltd, 2002.

11.2 “WindFarmer User Manual”, GH & Partners Ltd, 2002.

11.3 Frandsen S, “Turbulence and turbulence-generated fatigue loading in wind turbine


clusters”, Riso report Riso-R-1188(EN), 2003.

11.4 Adams, B.M., et al., “Dynamic loads in wind farms II,” Final report of CEC Joule
project JOU2-CT92-0094, GH report 286/R/1, 1996.

You might also like