AWFC Cordo
AWFC Cordo
AWFC Cordo
Relatore Candidato
Prof. Alberto Zasso Federico Cordò
Matr. 854532
Correlatore
Ing. Paolo Schito
Politecnico di Milano:
www.polimi.it
Scuola di Ingegneria Industriale e dell’Informazione:
www.ingindinf.polimi.it
Dipartimento di Meccanica:
www.mecc.polimi.it/
Contents
1 Introduction 1
1.1 Wind farm modelling . . . . . . . . . . . . . . . . . . . . . 3
1.2 Active wake control . . . . . . . . . . . . . . . . . . . . . . 7
1.3 Thesis outline . . . . . . . . . . . . . . . . . . . . . . . . . 10
3 Wake models 27
3.1 Bastankhah and Porte-Agel gaussian model . . . . . . . . 29
3.1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . 29
3.1.2 Experimental validation . . . . . . . . . . . . . . . 33
3.2 FLORIS: a control-oriented tool . . . . . . . . . . . . . . . 40
3.2.1 Additional code implementations . . . . . . . . . . 41
3.2.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.3 WFSim: a 2D RANS solver for control design . . . . . . . 49
3.3.1 Equations and computational framework . . . . . . 50
3.3.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.3.3 New turbine and mixing length model . . . . . . . 56
v
vi CONTENTS
Bibliography 97
List of Figures
vii
viii LIST OF FIGURES
ix
Abstract
Wind turbines are clustered in wind farms for economic reasons. Aero-
dynamic interaction between the wind and the turbine perturbates the
flow field, generating a flow structure called turbine wake. The wake is
characterized by increased incident turbulence and velocity deficit with
respect to the unperturbed velocity of the flow. Wakes interaction with
downstream turbines results in a decreased turbine lifetime and plant
efficiciency.
Nowadays these effects are neglected by the commercial wind farm
controller on the market. Each turbine is considered at individual level
and not as a player of a global system. Supercontroller capable to consider
aerodynamic interactions are being developed and allows to improve the
wind farm performances considerably.
In this thesis, wind farm control-oriented wake models are analysed.
Validation is carried out through experimental measurements performed at
the Polimi wind tunnel. This work was possible thanks to scaled G1 models
supplied by TU Munchen. The models under investigation are dynamic
as WindFarmSimulator (WFSim), and static as FLOw Redirection and
Induction in Steady-state (FLORIS), developed by TU Delft. Besides of
the validation procedure, model parameters and complexity are evaluated,
with a view to computational cost problems.
A proposal of supercontroller is developed with all components, indi-
vidually tested. Starting from standalone reference wind turbine controller,
supercontroller is introduced comprehensive of wake model and observer
for estimating all input parameters necessary to determine the control
action properly. Technique chosen to optimize the wind farm efficiency is
xi
xii LIST OF TABLES
xiii
xiv LIST OF TABLES
Introduction
1
2 Chapter 1. Introduction
Figure 1.1: Clouds forming in the Horns Rev offshore wind plant in Denmark
allows to visualize typical flow structure within the wind farm –
Photo
c Christian Steiness
If on one hand power losses affect the plant efficiency, on the other
asymmetric loads affect lifetime of downstream turbines. Currently, greedy
controls operate at an individual level, non-exploiting the overall wind
energy and, therefore, working in a non-optimal way. Wind farm control
algorithm needs wake models to estimate the wind flow and predict the
performances of the overall wind turbines. For open-loop control algorithm,
wake models are essential and the parameters has to be estimated properly
to reconstruct correctly wind flow while the closed-loop ones are based on
the measurements available within the wind farm. If measurements are
available, the wake model parameters may be online estimated to guarantee
4 Chapter 1. Introduction
good matching between the model suitable for the control design and the
real wind flow.
Figure 1.2: Development of the vertical profile shear in the three regions
ground and the shear of the ambient flow. In this regard, far wake region
is particularly suitable to be modelled as a gaussian profile[0].
Wind tunnel experiments are rather different with respect to full-scale
experiments because of wake meandering[0]. The importance of wake
meandering effect is pointed out by Larsen[0]. The wake meandering
causes continuous change of the lateral and longitudinal position of the
wake centre. It helps the overall efficiency of the plant, misaligning the
turbines most of the time. Since the scale effect and the influence of
terrain are assumed to be responsible for wake meandering, it should
be modelled as stochastic effect. To take into account this effect more
sophisticated numerical models re required, such as Dynamic Dynamic
Wake Meandering (DWM)[0] or Large Eddy Simulation (LES). However
designing a control exploiting this models is computationally expensive
and, nowadays, unfeasible.
Since many measurements from wind farm are available, at the actual
state of the art, second solution is preferable. Therefore, in this work,
only simple models are considered as FLOw Redirection and Induction
in Steady-state (FLORIS) and Wind Farm Simulation (WFSim). Both
were developed in TUDelft. FLORIS[0] is based on Basthakhan Porte-
Agel Gaussian model[0] and WFSim[0] is a 2-D unsteady RANS solver.
These models are respectively low and medium fidelity, FLORIS is static,
instead of WFSim, which is dynamic. The models are validated through
experimental wind tunnel tests in Chapter 3.
1.2. Active wake control 7
• Yaw control rotating the rotor and nacelle around the tower. In
this way it is possible to misalign the rotor with wind direction and
deflected the wake in the opposite direction to the yaw angle (γ).
• Tilt control rotating the rotor and nacelle in the tilt direction. In
this case the flow is redirected upward and downward in the opposite
direction to the tilt angle (τ ).
optimal TSR are: CT ≈ cos2 (γ) and CP ≈ cos3 (γ). Therefore, optimal yaw
configuration is a result of a compromise between partializing upstream
wind turbine and optimizing wake flow of the downstream wind turbine.
High fidelity simulations have shown that yaw control can increase power
of 4 − 7%[0], even if the topic is still actively reasearched.
13
14 Chapter 2. Wind turbines performances in yawed operations
treated correctly, but skew flow and dynamic stall are neglected. From
the control point of view, the model appears appropriate and code is
run quickly. Compromise between accuracy and low computational cost
is reached thanks to this model. Robust solution algorithm for BEM
theory is developed following AeroDyn procedure[0]. Incoming flow to
wind turbine are reconstructed according to Atmospheric Boundary Layer
(ABL) measured in wind tunnel test.
where:
• α is azimuthal angle
• γ is yaw angle
ξ = γ(0.6a + 1) (2.4)
mesh is not trivial, but in this way the distribution is assumed to be more
consistent with real wake induction. Drawbacks are represented by the
computational costs which increases and by the data of blades. Chord
and aerodynamic coefficients are needed to compute axial and tangential
coefficient. Aerodynamic coefficient are estimated from static wind tunnel
tests and are function of CD = CD (α, Re, r), CL = CL (α, Re, r). Therefore
a preliminary study of the airfoils has to be performed.
Figure 2.2: First CL-Windcon campaign: out- and in-wake flow characteriza-
tion allowing for different yaw misalignment angles and different
turbine power set-points.
2.2. Experimental tests 17
2.2.1 Setup
Regarding CL-windcon project, two specific configurations were tested.
The atmospheric boundary layer profile is generated using bricks and
spires[0]. Onshore configuration provides for turbulence around to 12%,
while offshore for 6%. Two configuration are fitted through power law
with α = 0.2 and α = 0.079 respectively. Two different models are tested,
but for this work, tests of G1 only are reported. G1 stays for generic
1-meter diameter rotor. G1 is conceived to satisfy several specific design
requirements:
Since many turbines are tested in wind tunnel to reproduce wind farm,
the turbines have to be very small because of blockage effect. Control is
also crucial for the testing of advanced control strategies. It is structured
in three distinct levels, thanks to Bachmann M1 system (modular real-time
controller) coupled to a CPU module for running control algorithm. Three
standard controls are implemented:
• Yaw control.
For BEM code validation in yawed operations, reference torque and collec-
tive pitch control adjust rotor speed depending on inflow condition. The
comparisons are carried out in comparable configurations, considering that
the TSR is not constant yawing the wind turbine.
With reference to Figure 2.3, measurements available for post-processing
are shown. A torque-meter, located after the two shaft bearings, allows for
18 Chapter 2. Wind turbines performances in yawed operations
W∞ (r)c(r)
Re(r) = (2.5)
q ν p p
W∞ (r) = Wx2 + Wy2 = V∞2 + (ωr)2 = V∞2 + (λr V∞ )2
p
= V∞ 1 + λ2r ≈ V∞ λr (2.6)
Figure 2.4: CT − λ and CP − λ are measured in some points and the curves
over the whole TSR domain are estimated through BEM-based fit.
First session In the first campaign, the working condition are set to
reference values: λ = 8.15, V∞ = 5.65m/s, blade pitch β = 1.4259◦ and yaw
angle over the range γ = {0◦ , 10◦ , 20◦ , 30◦ , 40◦ }. The tests are performed
2.2. Experimental tests 21
Figure 2.5: July Session: CT in yawed condition for both onshore and offshore
configurations; pitch angle, β = 1.4259◦ .
Figure 2.6: July Session: CP in yawed condition for both onshore and offshore
configurations; pitch angle, β = 1.4259◦ .
coefficient is best fitted through least square method, giving final value
of 15π
54
. Numerical results show good agreement with experimental tests.
Concerning low turbulence configuration, the wind profile input is modelled
as power law ABL, with an exponent α = 0.079. In this case, the maximum
error is lower than 3%. The model is capable to predict the performances of
the wind turbine, therefore suitable for control design. Same procedure is
followed for high turbulence configuration, fitting with exponential α = 0.2.
Even if the tests are performed at below rated, with increasing turbulence,
the λ decreases too. This effect is caused by torque control, which is
non-linear, since it is related to dynamic pressure and therefore to the
velocity square. Greater velocity contribution are amplified, resuting in
lower λ in high turbulence configuration. In this case, numerical model
shows error up to 9% for high yaw angle γ. The performance estimated by
P&P BEM are always better than the experimental. This fact might be
due to a worsening of the blades aerodynamic efficiency in high turbulence
wind profile. Nevertheless, for low angle yawed operation, the error is
below 5%. Therefore, in high turbulence boundary layer, finding optimal
set-point with a view to wind farm might be difficult.
Second session In the second campaign, the working condition are set to
λ = 8.15, V∞ = 5.75m/s, blade pitch β = 0.4259◦ . Lower pitch guarantees
to improve the match to theoretical scaled thrust force. In fact, both,
thrust force and torque at the rotor are enhanced by this modification. In
this test session, two turbines are aligned to estimate thrust and power
production of the downstream turbines also. Hereafter the results of the
upstream turbine are presented with a two-case-variability of the yaw angle
γ. Same considerations regarding control are still valid and processing is
carried out at the same way as July session. For both configurations errors
are lower than 5%.
The model is tested over wide range of pitch and yaw angle. In general,
good match with experimental results is achieved, paving the way of
an open-loop model-based control design. With increasing turbulence,
P&P BEM results get worse, even if the model seems to be capable to
2.2. Experimental tests 23
Figure 2.7: September Session: CT in yawed condition for both onshore and
offshore configurations; pitch angle, β = 0.4259◦ .
Figure 2.8: September Session: CP in yawed condition for both onshore and
offshore configurations; pitch angle, β = 0.4259◦ .
Results are similar and errors are always below 5%. In particular,
CP and power production is well estimated, while CT and thrust force
are slightly different. Wrong estimation of thrust has an impact in wake
reconstruction in FLORIS and WFSim. In fact, as shown in Chapter
2.3. Numerical validation of P&P BEM 25
3, thrust plays key role for estimating wake potential core and deficit.
Important fact to notice is that optimal λ is function of yaw angle γ. For
this reason, λ and γ are not two independant degree of freedom suitable
for control design, but λ depends on γ. Deeper analysis is presented in
Chapter 4, where it is shown that not adjusting λ in yawed operation,
up to 1% of the upstream wind turbine might be lost in terms of power
production.
Chapter 3
Wake models
The core of wind farm modelling is the wake model. The wake is
a flow structure that is characterized by a reduced wind speed and an
increased turbulence because of the turbine interactions with the incoming
flow. The problem depends on many variables both deterministic as the
morphology of the terrain and stochastic as time-varying atmospheric
conditions. Operating point of the wind turbine should adapt to these
conditions to work always in optimal way. Wake interaction effects have
been studied widely by Vermeer[0]. To catch all these aspects of the complex
physics behind wind turbine interaction, the problem investigation requires
sophisticated tools as CFD simulations. The accuracy of models, like
SOWFA, is very high, while they are very computationally expensive.
Therefore, nowadays, these models are not suitable for control design, but
allows to assess simple wake model. Several wake models are reported in
Table 1.1. In this chapter FLORIS and WFSim are analysed in detail in
terms of accuracy and computational cost.
In this work, active wake control is performed by means of yaw control.
Turbines working in yawed conditions have recently received considerable
attention both as: an unfavourable practical issue in the operation of wind
turbines and more importantly, as a favourable method to increase the
power production of whole wind farm. Indeed, wake deflection combined
to faster wake recovery allows to improve efficiency of the downstream
27
28 Chapter 3. Wake models
wind turbines[0] thanks to lateral mean momentum flux, which boosts the
flow entrainment in only one side of the wake due to yaw misalignment.
Therefore, the wake has non-symmetric distribution, with higher energy.
Models are expected to catch this fact and validation in yawed condition
is carried out by means of experimental wind tunnel data.
P&P BEM supplies the performances of the wind turbines to the two
models considered. FLORIS is a heuristic model and lacks a complete
near-wake model, while WFSim performs bidimensional RANS simulations.
Near-wake region is much more complex than far-wake region and it is
influenced highly by several factors such as tip speed ratio and wake rota-
tion. Describing near-wake region is useless and expensive effort for wind
farm control, since the gap among the turbines of the plant is greater than
the near-wake region. Therefore, downstream turbines always work in the
far-wake region of the upstream turbine. ABL and wind turbines working
conditions affect the wake severely. On one hand these aspects are properly
reproduced by experimental test, on other hand the models should be ca-
pable to consider all these parameters. To simplify the problem and reduce
computational costs, tuning variables are introduced. Further development
will allow to compute online these parameters to guarantee good matching
with real-time wind conditions. FLORIS is based on Bastankhah and
Porte-Agel gaussian (BPG) wake model[0]. With respect other models, like
Jensen[0], Ishiara[0] or Frandsen[0], BPG seems to be more solid and capa-
ble to reconstruct the wind turbine wake very effectively, even if keeping
high parsimony and low computational cost. As shown by Annoni[0], BPG
provides the best representation of the wake characteristics under different
atmospheric conditions, specifically accounting for turbulence intensity,
and different turbine operating conditions. The procedure is carried out as
suggested by Bastankhah[0] in the next section to assess the BPG model,
using G1 wind tunnel measurements in ABL. Model parameters are deeply
analysed and some modifications are proposed to improve the model. On
the other side, some WFSim simulations are performed and compared to
wind tunnel data, however from this analysis FLORIS seems much easier
to manage and provides better results tham WFSim.
3.1. Bastankhah and Porte-Agel gaussian model 29
3.1.1 Overview
The PIV wake characterization shows that self-similar gaussian distri-
bution represents good solution to fit the wake satisfactorily. However, for
yaw angle approaching to 30◦ , gaussian profile distorts the wake shape.
Fortunately, yaw angle equal or greater than 30◦ are rather impractical for
yaw-angle control methods, main purpose of this work. In fact, as already
said, for very high yaw angle, the upstream turbine undergoes to a too
high power decrement, such that the downstream one is unable to compen-
sate these losses. Moreover, self-similarity provides simplicity and general
validity to the model. Bluff-body wakes achieve self-similarity at a certain
downwind distance, i.e. every velocity deficit collapse onto a single curve,
changing the reference frame to dimensionless one through Buckingham
theorem. At the same time, these considerations are acceptable for wind
turbine wakes also, despite of wake deflection. In fact, concerning wind
30 Chapter 3. Wake models
Referring to Figure 3.1, the equation, which describe velocity and skew
angle can be written as:
−(y−δ) 2 2
−(z−zh )
ū(x, y, z) 2
= 1 − Ce 2σy e 2σz2 (3.1)
ū∞
−(y−δ+σy )2 −(z−z )2
ϑ(x, y, z) 2
h
= e 2σy e 2σz2 (3.2)
ϑm
where C is the velocity deficit at the wake centre normalized with the
incoming velocity and δ is the wake-centre deflection at each downwind
location. The h¯· i notation stays for time-averaged quantities. The wake
centre is assumed to remain at hub height zh as the vertical displacement
of the wake centre is rather small for lower yaw angles. C, σy , σz and δ
have to be determined to close the model.
The integral form of the RANS equation in the streamwise direction
has been extensively used in classical studies of bluff-body wakes[0]. Under
some assumptions as:
σy x − x0 σy0
= ky + (3.4)
d d d
σz x − x0 σ z 0
= kz + (3.5)
d d d
Where ky and kz are the wake growth rates, x0 is the onset distance of
the far-wake region and d is the turbine diameter. ky and kz are model
parameters and might be determined thanks to experimental tests or high
accurate numerical simulations. All model parameters are discussed in
the next section. Coupling actuator disk theory, Burton relation(Equation
2.4) for the skew angle and the streamwise momentum deficit allows to
compute the onset:
r
σ z0 1 σy0 σz
= , = 0 cos(γ) (3.6)
d 8 d d
the incoming turbulence also has an impact, enhancing the flow entrainment
and consequently the growth of the shear layer. Therefore, shear layer
width variation may consider both turbulence velocity gradient:
1 ds us ds ue
= = αI + β (3.7)
u∞ dt u∞ dx u∞
According to this equation, it can be readily shown that the length of the
potential core decreases with increasing thrust coefficient of the turbine,
incoming turbulence and yaw angle. These facts provide flow entrainment,
but while last two decrease wake deficit, first one has the opposite effect.
Therefore, increasing the thrust, the overall effect is greater subtraction of
energy from the flow, with shorter near-wake region characterized by high
velocity difference between potential core and unperturbed flow. Since x0
is strictly connected to the beginning of the far-wake region, tip-speed-ratio
and wake rotation have no contribution[0].
Once again, plug ū (Equation 3.1) and ϑ (Equation 3.2) into Equation
3.1. Bastankhah and Porte-Agel gaussian model 33
3.6, complex expression for δ is derived. Also in this case, the relation
presents a term due to onset of the far-wake region plus a upper bounded
logarithmic term, as x approaches to infinity.
x0
δ0 = ϑ0 (3.10)
d
wind tunnel mapping, periodically executed in GVPM. Wake data used for
the subsequent analysis are smoothed. Smoothing has minimum impact
on low turbulence measurements, while high turbulence wake is deeply
changed by this process. However, in this way, some notable features are
guaranteed like symmetry and regular wake shape, aspects that, otherwise,
BPG model would be not able to properly reconstruct. As a first impression,
from these figures it has to be noticed that deflection is strongly non-linear
and how much turbulence intensity enhance flow entrainment and therefore,
wake recovery.
Starting from these data, the potential of this model is shown. Pre-
3.1. Bastankhah and Porte-Agel gaussian model 35
where σy is the wake width in y direction, ūc is the velocity of the wake
centre and ŷ is the integration variable. The integral can be calculated
numerically to find the wake width at each downstream position. Note that
for a pure Gaussian profile reduces to the standard deviation of the profile.
As reported in Figure 3.4 self-similarity can be reasonably assumed for
lateral velocity proiles in the far wake. All wake measurements plotted with
respect wake width, maximum wake deficit and wake centre collapse onto a
single curve. This data sample is composed of thirty different configurations.
Even if many cases are considered, the coefficient of determination is close
to 1. High turbulent and high yaw angle data are the most difficult to
handle. In this regard, post-processing allows to smooth data and increase
quality for comparison. Much greater differences appears at the lateral
sides of wake. Fortunately, lateral sides are less important for active wake
control, since wake recovery is higher in these regions. In this case, it
is much more difficult to show precise differences between the two sides,
as reported by Bastankhah[0] in his studies. Indeed, theorically left side
should be influenced by strong spanwise velocity, resulting in less degree
of similarity.
36 Chapter 3. Wake models
growth rate and the shear layer variation rate with respect to turbulence
intensity and velocity difference.
Figure 3.8: Wake width as function of the yaw angle in the two turbulent
configurations
40 Chapter 3. Wake models
of FLORIS, but turbine model and wake interactions are mandatory for
wind farm simulations. Simple merging wake is required in FLORIS. Wind
turbine and BPG models have been already analyzed in detail, while wake
interactions have not been treated yet. Many wake interaction models are
available from literature[0]. Among engineering models, quadratic rule
from Katic[0] was validated and gave better agreement then simple linear
superposition. Local wakes are superposed to estimate the overall speed
deficit at the n-th downwind turbines, using a quadratic sum of square of
the local speed deficits:
v
u n
uX
∆un+1 =t ∆u2 |n+1i (3.12)
i=1
effect[0]:
−0.32
0.73 0.35 x
Iadded = Aoverlap 0.8ai I0 (3.13)
Di
where I0 is the ambient turbulence intensity and a is the axial induction
factor of the turbine. Then also turbulence intensities are summed up
thanks to Katic formula. This formulation was derived experimentally in
case of one single turbine. In wind farm scenario properly formulation
might be rather different from this one. Studies have shown that the added
turbulence intensity reaches an equilibrium point between two and three
turbines downstream[0] and decreases upon a certain distance, which in
FLORIS is fixed to 15D.
At last, optimization module is provided to compute optimal working
conditions for the wind farm. Control parameters are axial-induction and
yaw. Interior point algorithm is used to solve this non-linear optimization
problem. Typically, the cost function is defined as the power production,
however with implementation of aeroelastic code, different cost functions
can be chosen. For instance, weighted cost function between maximizing
power and minimizing flapwise and edgewise bending moment is preferable
to increase the wind turbines lifecycle.
CT = CT cospt γ (3.14)
CP = CP cospp γ (3.15)
pt ans pp are two tuning parameters that matches the power losses due to
yaw misalignment. As reported by Fleming[0], pp should be between 1.4
and 2.2. Jonkman found a pp of 1.88 for the NREL 5MW, performing LES
42 Chapter 3. Wake models
1 p
a≈ (1 − 1 − CT0 cosγ) (3.16)
2cosγ
Even though the model suggested by Glauert[0] can estimate the thrust
coefficient correctly, it is unable to properly evaluate the skew angle of
the flow through a yawed rotor. Reason accounts for this problem can
be found in the fact that this model only considers the component of the
induced velocity normal to the rotor, and not the tangential one. As a
result, the model is expected to overestimate the flow skew angle at the
rotor. For these reasons, Burton[0] formulation is preferable to estimate
skew angle:
44 Chapter 3. Wake models
ϑ = 0.6a (3.17)
0.3γ p
ϑ≈ (1 − 1 − CT0 cosγ) (3.18)
cosγ
In standard Floris code Equation 3.16 is plug into Equation 3.17, but
thanks to P&P BEM code, axial induction coefficient is directly computed.
Theorically, this method allows a more precise prediction of flow angle and
therefore, of wake deflection (Equation 3.10).
Jimenez[0] conjecture states that the wake growth rate increases with
increase in yaw angle. In incoming turbulent flow, the wake quickly
asymptotes to a turbulent free shear flow in the far-wake region. Therefore,
Bastankhah proposes that wake growth rate is not affected by the yaw angle,
as turbulence is the main mechanism responsible for flow entrainment. As
said, measurements of wake growth rate for the G1 model are in contrast
with both Bastankhah and Jimenez. Figure 3.11 shows precise trend with
the yaw angle, with low dispersion especially concerning low turbulence
case. As the turbulence increase, considerations by Bastankhah become
more relevant and yaw angle effect becomes negligible.
Figure 3.11: Wake growth rate as function of the yaw angle in the two turbulent
configuration
Linear trend can reasonably fit the yaw angle trend as shown by coeffi-
cient of determinaton. Therefore, different formulation for wake growth
rate is introduced in the code to improve agreement with experimental
3.2. FLORIS: a control-oriented tool 45
data:
I−Iref
1 Iref
k = ka I + kb γ + kc (3.19)
10
ka = .124, kb = .00088, kc = .0236, Iref = .05
3.2.2 Results
All tested configurations in GVPM are simulated by FLORIS. Wake
growth rate is set up as explained in previous section. The remaining
tuning variables, α and β (Equation 3.8), are chosen to minimize error
between measured and numerical wakes. Proper tuning allows to achieve
good prediction of the wake features. For the sake of completeness, two
sets of measurements should be performed: the first one useful to identify
model parameter and the second to validate the model. In this case,
identifying set only is available. In this phase, a preliminary analysis of the
model is performed, since the ultimate purpose of this work is predicting
optimal operating condition. Thrust and Power are computed by P&P
BEM code. Turbine parameters, like TSR, pitch and rotor speed are
consistent with the time averaged experimental data. Both loads and wake
measurements were performed in July session. Therefore, measurement
set-up and control logic are the same explained in Chapter 2.2. Probes are
equally spaced by 100mm one from each others, providing up to twelve
46 Chapter 3. Wake models
Figure 3.12: Wake data: FLORIS and experimental at low TI for different
yaw angle configurations {0◦ , 10◦ , 20◦ , 30◦ }
In Figure 3.12, results in low turbulence shear flow are presented. For
low angle misalignment, results show very good agreement with experi-
mental tests. Nonetheless, concerning all configurations, error is always
below 5%. At larger downstream locations, the model matches better the
experimental data, since wake deficit is lower and flow entrainment has
already reintroduced and distributed energy to the flow within turbine
diameter. Maximum wake deficit is properly reconstructed while wake
centre is slightly different with respect to experimental tests (Figure 3.14).
3.2. FLORIS: a control-oriented tool 47
Figure 3.13: Wake data: FLORIS and experimental at high TI for different
yaw angle configuration {0◦ , 10◦ , 20◦ , 30◦ }
In Figure 3.13, high turbulence results are shown. Also in this case, as
reported in Figure 3.15, wake deficit is reproduced with good approximation,
48 Chapter 3. Wake models
but wake centre position is not properly caught by the model. Increasing
yaw misalignment, the spaced-averaged wake velocity is slightly different
from the experimental one and the error approaches to 5%, which is
reasonable estimation error.
As any wind farm model, wake, turbines and turbulence model are
needed to define the problems completely. Each turbine is modelled
through actuator disk model (ADM) and 2D Navier-Stokes are used to
derive a model with a heuristic turbulence feature. In this way, the
resulting matrices from discretization process are sparse, leading to low
computational effort.
As starting point, 3D incompressible filtered Navier-Stokes equation
are used, as in LES case:
∂ṽ 1
+ ṽ · ∇ṽ + ∇ · τm + ∇p̃ − f = 0 (3.20)
∂t %
∇ · ṽ = 0 (3.21)
be derived:
∂w
∇·u=− (3.23)
∂z
With u = (ṽ1 (x, y, zh ), ṽ2 (x, y, zh ))T , τH is the 2D stress tensor and
e1 , e2 are the two unit vector along x and y. The right-hand side is
roughly neglected in the Equation 3.22 and to adjust continuity equation
it is supposed that for a quasi axis-symmetric case: ∂w ∂z
∂v
≈ ∂y . This last
additional term allows to avoid senseless speed-up in the flow. Therefore,
the set of equation which has to be discretized is rather similar to RANS
equation, even if model is derived starting from LES:
∂u
+ u · ∇u + ∇ · τH + ∇p − f = 0 (3.24)
∂t
∂v
∇·u=− (3.25)
∂y
The set of equations are spatial discretized over a staggered grid follow-
ing Versteeg[0]. It is carried out by employing the Finite Volume Method
and the Hybrid Differencing scheme. Temporal discretization is performed
using the implicit method that is unconditionally stable. Dirichlet inflow
boundary condition are imposed to one boundary, while on the others
Neumeann (zero stress) conditions are prescribed. Initial pressure field is
set to zero.
The most controversial issue concerning the model is the mixing length
turbulence definition. Of course, mixing length relates eddy viscosity to
∂u
∂y
, following Prandtl[0].
2 ∂u
τh = −νt S; νt = lu (x, y) (3.26)
∂y
Mapping mixing length over the entire domain is a difficult effort.
However, the terms related to mixing length enter in x-momentum and y-
52 Chapter 3. Wake models
(x0 − n0 )l , if x0n ∈ Xn0 and if yn0 ∈ Yn0
n s
lun (x0n , yn0 ) = (3.27)
0, otherwise
With the notation h ·0n i, n-th turbine local system is indicated. Xn0 =
{x0n m0 ≤ x0n ≥ n0 } and. Yn0 = {yn0 |yn0 |0 ≤ D}, where m’ represents the
near wake region limit, while n’ defines the region affected by the wind
turbine interaction. Providing this formulation only three parameters have
to be estimated, but it might occurr that the formulation is not be able
to capture all the complexity of the wake physics in this way. Moreover,
m’ can be estimated a fortiori, leading to very simple formulation suitable
for online estimation. ls is the main tuning variable and expresses the
amount of wake recovery. According to Lee[0], as said in previous section,
ls should be influenced by ambient turbulence for sure, but also by turbine
characteristics as thrust coefficient and yaw angle. This formulation might
be not valid in case of two superimposed wake. Indeed, in FLORIS the two
wakes, and, therefore, the wake growth rates are considered separately, then
summed up following Katic formula, while in this case the two contributions
are considered together in the x-momentum and y-momentum. In case of
small wind farm, where several wakes interact each others, this fact might
lead to estimation error.
At last, turbine model is provided to estimate wind turbine perfor-
mances thanks to classical non-rotating actuator disk. Thrust force is
uniformly distributed on the wind turbine mesh cells. This is a critical
point, since this distribution is not consistent with the problem physics.
The distribution of thrust force has greater impact in near-wake region
than in the far-wake one. Therefore, estimation errors are expected in
3.3. WFSim: a 2D RANS solver for control design 53
2
1 2 1 Urp
T = %ACT U∞ = %ACT
2 2 1−a
1 CT 1
= %A 2
2
Urp = %ACT0 Urp
2
(3.28)
2 (1 − a) 2
where Urp is the flow velocity at the rotor plane computed following
actuator disk theory. Of course, induction is not considered directly, but it
derives from the numerical computations. Furthermore, a tuning parameter
is introduced to match the thrust force and flow velocity at the rotor plane,
which can be assumed to be independent of mixing length turbulence.
At the same way, the power is calculated from the resolved flow velocity
components. As already said, both CP and CT are input parameters. For
these reasons, WFSim performs at its own best, if coupled to high fidelity
software, like SOWFA. In this way, as reported by Doekemeijer, tuning
variables and performances coefficients can be estimated thanks to Kalman
filter.
3.3.2 Results
Mesh Time
L[m] N Points per D Aspect Ratio Tend [s] dT [s] Courant
x 34 50 1.617 1:4.87 10 0.1 0.822
y 14 100 7.857
Table 3.1: WFSim simulation parameters
be fixed changing the thrust distribution only. Results are very sensitive to
mesh grid and, therefore, the problem might be detected in cells dimension.
However, restriction on computational time limits the size of the cells.
Of course, changing the distribution of the thrust force, mixing length is
changed also over the x direction. From this preliminary analysis, WFSim
is reported to be not suitable for yaw control design in this configuration.
Since main problem of WFSim are related to wake deflection, it might
be useful for axial induction control. However, concerning yaw control, it
needs to be coupled to a high fidelity model or wind farm real measurements
to estimate online model parameters.
Figure 3.16: Wake data: WFSim and experimental at high TI for different
yaw angle configuration {0◦ , 10◦ , 20◦ , 30◦ }
56 Chapter 3. Wake models
Figure 3.17: Axial force distribution and blade factor, implemented in WFSim.
Turbine parameters are the same as in September session λ = 8.15,
β = 0.4259, γ = 0◦
Iungo[0] and Madsen[0] state that mixing length can be assumed constant
along the y direction. In this case, the wake is distorted in the centre, due
to impractical zero axial thrust force in that zone. New parameter, Blade
factor is introduced and reasonably fixed at 30% of the blade. Within
3.3. WFSim: a 2D RANS solver for control design 57
the blade factor, axial force is kept constant to the value of the axial
force evaluated at the blade factor. To maintain overall axial force, the
distribution is scaled. In this way, the axial force integrated over the rotor
diameter is not changed.
According to these significant additional changes to turbine-flow inter-
actions, new mixing length is introduced. Adding some degrees of freedom
to mixing length accounts for better matching with experimental results.
On one hand the new formulation should be effective in reproducing ex-
perimental results, on other hand mixing length has still physical meaning
and it cannot be turned upside down with respect to the original one. In
this regard, a good proposal is formulated by Madsen[0]. As said, wake
recover is affected mainly by ambient turbulence and slightly by turbines
parameter. Original mixing length in WFSim is responsible for ambient
turbulence only and it is assumed to vary linear over the x direction.
According to eddy viscosity concept, in this case slightly different with
respect to Equation 3.26:
νt = lm Umi (3.29)
Where lm and Umi are suitable turbulence length scale and velocity. νt
provided by Madsen is chosen to be dependant of ambient turbulence and
wake shear layer:
by means of filter functions, which allow to decouple and weight the two
effects separately. F1 filter function varies linearly from 0 to 1 over the
distance 0 to 2D, while F2 varies quadratically up to 10D.
Furthermore, in y direction smooth trend is chosen from the tip of rotor
diameter to one additional external diameter. This solution has physical
meaning and decreases speed up at the lateral side. New resulting mixing
length mapping is shown in Figure 3.18.
Figure 3.18: Mixing length turbulence new mapping, according to P&P BEM
new axial force distribution
Figure 3.19: Wake data: WFSim with new mixing length and turbine model
and experimental at high TI for different yaw angle configuration
{0◦ , 10◦ , 20◦ , 30◦ }
Chapter 4
61
Chapter 4. From standard controller through wind farm
62 algorithm
1 CP,opt
k = %πR5 3 (4.2)
2 λ
Figure 4.3: Rotor speed response to pitch step input: Equlibrium position is
perturbated at different inflow condition. Transfer function are
estimated by this process
between saturated signal and the original one provides anti-windup term.
By trial and error procedure, the gain is set to 1000rad/s/rad. In this
way, if the pitch is not saturated, anti-windup signal is zero, while, if the
pitch is saturated, the error is not integrated, and, therefore a delay in
control action is avoided. Both coupled torque and collective pitch control
defines the reference control for wind turbines.
4.1.3 Simulations
Numerical analysis is carried out by means of FAST. This analysis
presents many approximations with respect to experimental tests. Lift and
Static coefficients are constant with Reynolds, the performances of the wind
turbine are reported in Figure 2.9. Elastic transmission is considered as
infinitely rigid. Thus, rotor speed and generator speed are perfectly equal.
Besides of these wind turbine related approximations, inflow conditions are
rather different. Indeed, the simulations are performed at uniform wind
speed. Shear and time history of the inflow wind are consistent with the
wind tunnel data, however the wind is uniform along the y direction.
These tests are not a check for the control design but aim to understand
the behaviour of the wind turbine in ambient turbulence. In Chapter 2, it
has illustrated that with increasing ambient turbulence, the performances
decrease and the control gets trouble to keep constant the TSR. In this
analysis, this fact cannot be seen, due to many approximation and, in
4.1. Reference torque and pitch controller 69
trated by Bottasso [0]. Since turbulence intensity level and wind direction
are input variables for both wake models, FLORIS and WFSim, they
are the reasons of interest of the observer. While wind direction can be
directly estimated by the observer, turbulence intensity can be derived by
the wind shear. Basic idea of the observer design is to exploit 1P response
of the measured flapwise bending moment. Starting from the physics of
the problem, general linear formulation is derived, whose parameters are
obtained by system identification, based on the least-square method.
4.2.1 Formulation
d2 β γ
4 dβ γ
2
+ 1 − cosψ(Ū0 + q̄l) + (K + 2Bcosψ + Ū0 sinψ)β =
dψ 8 3 dψ 6
γ q̄
= (A − sinψ) − 2q̄cosψ−
2 4
γ λ 2ϑp K1 V̄0
− (Ū0 + q̄l) + + cosψ (4.5)
2 2 3 4
Plug the Equation 4.6 into the Equation 4.5 and rearranging the terms, it
is possible to invert the system and compute cross-flow and vertical wind
4.2. Observer for estimating wind direction and shear 73
shear as:
" # " #
Ū0 0
= +
K1 V̄0 −16Bβ0 /γ
" #" #
3/4 −6(K − 1)/γ β1c /β0
+ +
−8(K − 1)β0 /γ − 3A3 −β0 + 24A3 (K − 1)/γ −β1s /β0
" #
−3/(4β0 )
+ q̄ (4.7)
−16/γ + A3 (3/β0 − 4l)
Since blade flap angle is typically not available from wind turbine
measurements, this formulation is impractical. However, supposing direct
relation between flapwise bending moment and blade flap angle, general
observation model can be introduced with similar one to the previous
one, which is validated by the physics of the problem. In this case, the
matrices coefficients are not linked to the physical quantities but are
identified using least-square procedure by identification tests set. The
general wind-scheduled observation model can be reformulated:
transform:
Z ψ−2N π
1
m0b = mb (t)dψ (4.12)
2N π ψ
Z ψ−2N π
1
m1cb = mb (t)cosψdψ (4.13)
Nπ ψ
Z ψ−2N π
1
m1sb = mb (t)sinψdψ (4.14)
Nπ ψ
4.2.2 Simulations
Ambient turbulent results are shown in Figure 4.10. Wind speed time
history and the shear power law of 0.079 are consistent with the data
measured by the pitot tube during wind tunnel test in low turbulence
configuration. Therefore, the inflow conditions are similar to real case,
despite of the perfect spatial coherence over the rotor area. Theoretically,
the coherence has a negative impact on the performances of the observer.
Real turbulence varying over the rotor area should balance itself, helping
disturbance rejection. Instead of previous case, this time gain scheduling
is involved.
4.13 on the left, the model can predict the operating point of the wind
turbine working in yawed condition with reasonable accuracy. Coupling
point is obtained by the superimposition of the aerodynamic torque and
of ideal generator torque fixing the working condition as TSR = 8.15. It
has to be noticed that as the yaw progressively increases, the aerodynamic
torque decreases and TSR decreases as well. The amount of the decrement
in terms of torque and TSR does not depend on this kind of torque control,
while it depends only on the aerodynamic characteristic of the wind turbine.
In practise, this torque control is not able to control the TSR, as the wind
turbine works in yawed condition. The curves are obtained fixing incoming
flow velocity U = 5.65 m/s as in the wind tunnel tests. It should be added
that no relation between TSR and normal flow can be obtained, since the
phenomenon is extremely non linear.
On the left of the Figure 4.13, optimal TSR is illustrated. The inter-
sections between the black curve and the coloured ones represent the real
working condition with standard torque control. Of course, optimal TSR
is computed by simple numerical approach as BEM and, therefore, it is
affected by estimation error. However, significant differences between the
two points are present, such that it can not be justified by numerical error.
Chapter 4. From standard controller through wind farm
84 algorithm
Yaw misalignment [◦ ]
40 30 20
Operating cond. Opt. 55.9-0.4465 60.98-0.5046 66.06-0.5458
ω[rad/s] - T[Nm] Std. 60-0.41 64.6-0.476 68.05-0.526
Opt. 24.95 30.77 36.05
Power [W]
Std. 24.6 30.74 35.8
Decrement -1.4% -0.1% -0.7%
Table 4.2: Torque control in yawed condition
4.3.2 Architecture
The architecture of wind farm controller is illustrated in Figure 4.14.
The control logic is based on active wake redirection thanks to yaw con-
trol. Each element is assessed by high fidelity numerical simulation or
experimental test. However, in this work, the structure is not tested as a
whole. Further developments might be addressed to supercontroller testing
within the SOWFA environment. In the last Chapter, FLORIS is used to
estimate optimal set-point of 3x1 turbines array. The results are compared
to experimental set-point found by means of trial and error procedure
changing progressively the yaw settings of wind turbines array.
The wind farm supercontroller architecture needs the wake model to
consider the overall wind farm, including mutual interaction among wind
turbine wakes, the observer to estimate state-space variables, which cannot
be directly measured due to practical issue and standard wind turbine con-
trollers. Wind farm can work as an active network, exchanging information
in real-time. In this particular case, wake model needs the data of incoming
flow, as mean wind speed, incoming turbulence and wind direction. The
mean speed is supposed to measured, while other two input variables are
4.3. A proposal for the supercontroller structure 85
Figure 4.14: Wind farm control architecture composed of wind observer, wind
fam optimizer and greedy controller.
Chapter 5
At the end, some conclusions and future outlooks are illustrated. Indeed,
from this study great potential of yaw control is revealed, but further steps
are required to improve FLORIS model and to develop real closed-loop
control architecture for wind farm operations.
87
88 Chapter 5. Wind turbines array optimization
5.1 Results
In the CL-windcon campaign of May 2018, 3x1 wind turbines array
configuration is tested in Polimi wind tunnel. Several tests are performed
varying the wind direction. In this analysis, four different cases are con-
sidered consistent with practical case. The yaw direction is obtained by
misaligning the wind turbines along the lateral direction:
The results are provided below. FLORIS model parameters are tuned
following the wake measurements of the first CL-windcon campaign. There-
fore, the tuning is not properly dedicated to this case. In addition, added
turbulence model parameters are not changed with respect to the original
ones. Although model parameters are unaltered, a scaling factor is intro-
duced to fit the experimental results with the numerical. This scaling factor
is constant and, therefore, the yaw trend is unaffected by this coefficient.
In practise, the optimal point is not changed by this corrective term, which
accounts for model errors, measurements uncertainty and wind tunnel
blockage effects.
The model shows good agreement with experiments for not perfectly
aligned configuration, while progressively changing the wind direction to
aligned wind turbines configuration, the experimental yaw trend becomes
rather different. In the perfectly aligned configuration, the numerical
yaw trend of the second turbine is totally in contrast with experimental
data. Probably, wake superimposition is over simplified in FLORIS model,
5.1. Results 89
Numerical Experimental
Baseline CP 1.226 1.229
Optimum 6◦ /6◦ 11◦ /6◦
Optimum CP 1.25 1.25
Efficiency +1.9% +1.7%
RMS 0.0082 /
5.1. Results 91
Numerical Experimental
Baseline CP 1.144 1.142
Optimum 13◦ /14◦ 15◦ /10◦
Optimum CP 1.2 1.204
Efficiency +5.2% +5.4%
RMS 0.0127 /
92 Chapter 5. Wind turbines array optimization
Numerical Experimental
Baseline CP 0.8392 0.826
Optimum 24◦ /23◦ 27◦ /15◦
Optimum CP 1.031 1.032
Efficiency +23.5% +24.8%
RMS 0.0197 /
5.1. Results 93
Numerical Experimental
Baseline CP 0.698 0.699
Optimum 23◦ /29◦ 31◦ /19◦
Optimum CP 0.737 0.746
Efficiency +5.6% +6.7%
RMS 0.0236 /
94 Chapter 5. Wind turbines array optimization
5.2 Conclusions
The basic idea of this work is to develop a simple control architecture
for wind farm to optimize the performances of the overall wind farm. Wind
farm numerical models are validated thanks to wind tunnel measurements.
The control architecture is closed-loop by means of the observer, but the
wake model is pre-tuned and, therefore, the computational cost is almost
non-existent.
First, the control-oriented wake models are investigated by means of
wind tunnel data in low and high turbulence ABL. BPG wake model is
validated and dedicated modifications are implemented according to wake
measurements. From the analyses, FLORIS results to be capable to prop-
erly reproduce the flow field within the wind farm in offline configurations,
while WFSim needs online tuning of model parameters to get adequate
results. For this reason, in wind tunnel testing, FLORIS is preferable to
estimate and predict optimal working condition for wind farm.
Secondly, simple formulation of control architecture for wind farm is
formulated. Starting from an overview of reference control, an observer
for estimating wind conditions, as wind shear and direction is developed.
The observer needs blade root bending moments to estimate the inflow
conditions, which are general available for commercial wind turbines. In
this sense, the control architecture can exploit wind farm information
as communicating network, in agreement with wind farm closed-loop
paradigm. The observer is effective in low turbulence configuration, but it
needs some corrections in the high turbulence atmospheric inflow. In this
way, it is possible to estimate input variable for the FLORIS wake model.
In last chapter, the real performances of FLORIS are tested in wind
tunnel environment for 3x1 wind turbines array in low turbulence ABL. For
array misaligned with wind direction, the FLORIS model is able to estimate
the optimum set-point with good agreement with respect to experimental
results, while for array aligned with wind direction, the model is in contrast
with experimental results.
In conclusion, this work defined as a preliminary analysis in wind farm
5.2. Conclusions 95
97
98 Bibliography
[0] Ewan Machefaux, Gunner Chr Larsen, and JP Murcia Leon. “Engineer-
ing models for merging wakes in wind farm optimization applications”.
In: Journal of Physics: Conference Series. Vol. 625. 1. IOP Publishing.
2015, p. 012037.
[0] I Katic, J Højstrup, and Niels Otto Jensen. “A simple model for
cluster efficiency”. In: European wind energy association conference
and exhibition. 1986, pp. 407–410.
[0] Amin Niayifar and Fernando Porté-Agel. “A new analytical model
for wind farm power prediction”. In: Journal of Physics: Conference
Series. Vol. 625. 1. IOP Publishing. 2015, p. 012039.
[0] Leonardo P Chamorro and Fernando Porte-Agel. “Turbulent flow
inside and above a wind farm: a wind-tunnel study”. In: Energies 4.11
(2011), pp. 1916–1936.
[0] Paul Fleming et al. “Field test of wake steering at an offshore wind
farm”. In: Wind Energy Science 2.1 (2017), p. 229.
[0] Hermann Glauert. A general theory of the autogyro. Vol. 1111. HM
Stationery Office, 1926.
[0] S Boersma et al. “A control-oriented dynamic wind farm flow model:“WFSim””.
In: Journal of Physics: Conference Series. Vol. 753. 3. IOP Publishing.
2016, p. 032005.
[0] Henk Kaarle Versteeg and Weeratunge Malalasekera. An introduction
to computational fluid dynamics: the finite volume method. Pearson
Education, 2007.
[0] Ludwig Prandtl. “Bericht uber Untersuchungen zur ausgebildeten
Turbulenz”. In: Zs. angew. Math. Mech. 5 (1925), pp. 136–139.
[0] Giacomo V Iungo et al. “Data-driven RANS for simulations of large
wind farms”. In: Journal of Physics: Conference Series. Vol. 625. 1.
IOP Publishing. 2015, p. 012025.
Bibliography 101