Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Applied Energy 303 (2021) 117691

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Formulas of the optimized yaw angles for cooperative control of wind farms
with aligned turbines to maximize the power production
Hongliang Ma , Mingwei Ge *, Guangxing Wu , Bowen Du , Yongqian Liu
State Key Laboratory of Alternate Electrical Power System with Renewable Energy Sources, North China Electric Power University, Beijing, China

H I G H L I G H T S

• Cooperative yaw control of aligned turbines is studied for power maximization.


• Two algebra formulas are proposed to determine the yaw angles for wind farm control.
• The formulas are efficient and generalized for different wind farm parameters.
• The proposed strategy is verified via LES and achieves significant power increase.

A R T I C L E I N F O A B S T R A C T

Keywords: Cooperative yaw control of wind turbines can substantially increase the total power production of wind farms,
Cooperative yaw control especially for the ones with aligned turbines. However, a generalized and effective method to quickly determine
Wind farms with aligned turbines the turbines’ yaw angles for the wind farm control is rather limited. To this end, a systematic study was carried
Power maximization
out on the cooperative yaw control of a single column of turbines. It is found that the optimized yaw angles are
Large-eddy simulation
not sensitive to the turbine number. All the turbines should yaw to the identical direction for power maximi­
zation except the last turbine for which no yaw is required. The cooperative yaw control strategy can be
simplified to a solution space only containing two yaw angles, i.e., the yaw angle of the first turbine and the angle
for the downstream yawed turbines. Then, two algebra formulas are proposed to quickly determine the two yaw
angles for the wind farm control only using three available parameters, i.e., the thrust coefficient of the rotor, the
wake decay coefficient and the turbine spacing. To validate the proposed formulas, high-fidelity large-eddy
simulations are performed. The results demonstrate that the proposed control strategy can increase the power
production significantly for cases with strong wake effects. For one of the cases with five aligned turbines, the
total power can be enhanced by 17.5%.

1. Introduction is of great significance for the operation of a wind farm.


A host of studies are performed to reduce the wake losses of wind
In large wind farms, the wake effects of wind turbines can reduce the farms. For wind farms in the planning stage, the micro-sitting optimi­
incoming wind speed and the operating efficiency of the downstream zation was often resorted to reduce the wake losses [3–4]. For this stage,
turbines. Through field measurement at Nørrekær, an onshore wind varieties of optimization algorithms [5–6], CFD methods [7–8], and
farm in Denmark [1], a more than 40% power loss is observed for the wake models [9–10] were proposed, and multiple micro-sitting software
downstream turbines due to the wake effect in the alignment direction. packages were developed, such as the Windsim, the Windfarmer, the
In addition, high turbulence intensity will be triggered by the upstream WT, etc. While for wind farms in service, the intelligent farm control to
turbines, which can substantially increase the fatigue loads of the tur­ increase the total power output were also extensively studied. Recently,
bines immersed in the wakes. Thomsen et al. [2] found that the fatigue the cooperative yaw control, which has been proven to promote the total
loads of turbines operating in wakes can be promoted by 5–15% due to power of a wind farm [11–13], shows a good prospect in engineering
the enhancement of the turbulence intensity. Therefore, the wake effect applications and hasbecome a research frontier in the field of wind

* Corresponding author.
E-mail address: gemingwei@ncepu.edu.cn (M. Ge).

https://doi.org/10.1016/j.apenergy.2021.117691
Received 13 February 2021; Received in revised form 14 August 2021; Accepted 22 August 2021
Available online 30 August 2021
0306-2619/© 2021 Elsevier Ltd. All rights reserved.
H. Ma et al. Applied Energy 303 (2021) 117691

energy. But in a realistic wind farm, the wake deflections are highly correlated
with the thrust coefficient of the rotors (CT), the wake decay coefficient
1.1. Progress of cooperative yaw control (kw), and the turbine spacing (Sx). The control strategies suitable for
various parameters mentioned above are not satisfactory, which hinders
Due to the interaction between the rotor and the incoming flow, the the development and engineering application of the cooperative control
wake of a yawed turbine deflects significantly in the crosswise direction. of wind farms.
Grant and parkin [14–15] found that the wake of a yawed turbine de­ To address the above issues, the cooperative yaw control of a single
flects opposite to the yaw direction based on wind tunnel measurements. column of turbines is studied systematically. Two algebra formulas are
Jiménez et al. [16] and Howland et al. [17] further studied the deflec­ proposed to quickly determine the yaw angles of the turbines for the
tion of a turbine wake at different yaw angles via both physical exper­ wind farm control to maximize the total power. In the construction of
iments and large-eddy simulation (LES). They found that when the the two formulas, the thrust coefficients, incoming turbulence in­
turbine operates at a yaw angle of 30◦ , the wake deflection at 7D tensities, and turbine spacings are considered. To validate the proposed
downstream (D is the rotor diameter) can reach 0.5D-0.7D. Based on the formulas, LES studies were carried out and the results show that the
wake deflection, preliminary studies were performed on the cooperative present strategy can substantially reduce the wake losses of the wind
yaw control for cases with two aligned turbines. Fleming et al. [18–19] farms with aligned turbines.
claimed that due to the wake deflection of the upstream yawed turbines, This paper is organized as follows. In Section 2, the optimization
the power generation of the downstream turbines is enhanced and the method of cooperative yaw control including the power calculation
fatigue loads of these turbines are substantially reduced. Adaramola and method and the optimization procedure is introduced. In Section 3, the
Krogstad [20] studied the yaw control of two model turbines aligned optimized results are shown and two algebra formulas to determine the
with a spacing of 4D through wind tunnel measurements, and it is turbines’ yaw angles are proposed with different thrust coefficients,
showed that when the upstream turbine yaws 30◦ and the two turbines wake decay coefficients, and turbine spacings. In Section 4, the proposed
stagger with a spanwise spacing of 0.4D, the shelter effect of the up­ formulas are validated numerically by high-fidelity LES. Finally, the
stream turbine on the downstream one is remarkably reduced, and a summary and conclusions are given in Section 5.
12% total power increment is achieved. Bartl et al. [21] investigated the
yaw control of two wind turbines for different turbulence intensities and 2. Optimization method of the cooperative yaw control of a
inflow directions through measurements, and their results showed that wind farm
the total power can be promoted by 3.5–11% through the cooperative
control. In this section, a wake model is selected to quickly compute the
In addition to the above studies, many researchers extended the shelter effect between yawed wind turbines in a wind farm. To this end,
cooperative yaw control to cases with more wind turbines. Taking four four wake models are compared with the results of high-fidelity LES.
aligned wind turbines as a research object, Park and Law [22] tried to According to the comparison, the Shapiro model [29] is adopted in this
optimize the total power of the turbines by a real-time cooperative study. Then, the yaw angles of the turbine are optimized to maximize the
control of the pitch and yaw angles using a data-driven method. Nu­ total power production of the wind farm using the genetic algorithm.
merical results showed that the method can increase the total power by
about 27%. Bastankhah and Porté-Agel [23–25] investigated the coop­ 2.1. Wake models for yawed turbines
erative yaw control for the case with five aligned turbines through wind
tunnel experiments. They found the total power can be added by 17% Wake models for yawed wind turbines can quickly predict the wake
using an optimized yaw control. Furthermore, Gebraad et al. [26] con­ of wind turbines under yawed conditions, and the accuracy of the
ducted the cooperative yaw control on a wind farm with 3 (streamwise) models was validated by many numerical simulations and experiments
× 2 (spanwise) wind turbines. A 13% increase of the total power is [11,30–32]. In contrast, the CFD approach is too computational
achieved, and in addition, the fatigue loads of the most of the turbines expensive to be used in a real-time assessment of a wind farm. Therefore,
are significantly reduced. Archer et al. [27] conducted the cooperative in this study, the wake models will be resorted to make a rapid predic­
control in a realistic wind farm with 28 wind turbines with LES method, tion of the wind farm’s power output. In the following, four candidate
and they found that positive yaw angles in the Northern Hemisphere can models are briefly introduced. (i) The pioneer model proposed by
enhance the total power generation due to the Coriolis force. The best Jiménez et al. (Jiménez model). The model is an extension of the
result is obtained with the yaw angle of 20◦ for the upstream turbines Frandsen model [33] accounting for the deflection of the wake under
and 10◦ for the deep-row ones. Recently, the cooperative yaw control yawed condition. In essence, this model is a quasi-one-dimensional
was investigated by Howland et al. [28] in a realistic wind farm with six model since the radial variation of the velocity deficit in the wake is
aligned wind turbines. The field measurement results shows that when neglected. The accuracy of this model was well validated by Jiménez
the inflow wind speeds are close to the average wind speed, the total et al. [16] with LES. (ii) The two-dimensional wake model proposed by
power increment can achieve 7–13%, while at night the increment can Bastankhah and Porté-Agel (BP model) [11]. The model is developed
even reach 28–47% under low wind speeds. according to the self-similarity property of the far wake and the concept
of the starting point of the far wake (i.e., the wake deflects linearly in the
1.2. Current shortcomings and the present study near wake region, and the mean velocity deficit conforms to the self-
similarity law behind the starting point of the far wake). (iii) The Qian
As described above, the total power of a wind farm can be remark­ and Ishihara’s model (QI model) [34]. In this model, both the effects of
ably enhanced through cooperative yaw control, but there still exist turbulence intensity and thrust coefficient are considered. The model
some shortcomings. (i) In previous studies, wind tunnel experiments or was well validated through high-fidelity LES by Qian and Ishihara. (iv)
numerical simulations with high computational or experimental cost The model proposed by Shapiro et al. [29]. In the model, the flow near
were mostly used, and most of the control strategies were determined the rotor is assumed as an inviscid region, and the yawed rotor is
empirically or by enumeration. It still lacks an efficient method to regarded as an elliptical-loaded lifting line. Based on these assumptions,
determine the yaw angles of the turbines in a realistic wind farm control. the counter vortex pairs and the initial spanwise velocity deficit can be
(ii) Most of the studies were only carried out under specific conditions. determined. Then the average velocity on the yawed rotor and the

2
H. Ma et al. Applied Energy 303 (2021) 117691

streamwise velocity deficit are calculated using the momentum theorem ∫∫


yδu(x, y, z)dydz
and Bernoulli equation. Compared with other models such as the yc (x) = ∫∫Ω (1)
Jiménez model and the BP model, the prediction accuracy of the wake Ω
δu(x, y, z)dydz
trajectory is improved [29].
where, δu is the velocity deficit of the turbine wake while Ω is the cross-
Before proceeding further, we would like to select one model from
section of the flow tube through the wind rotor. Fig. 5 illustrates the
the above four wake models according to the LES results. To generate
comparison between the wake centerlines from the wake models and the
high-quality LES data, the in-house LES solver, LESGO, is used to
perform the simulations. The solver was validated by experiments and LES results. Obviously, the Jiménez model overestimates the wake
has been widely adopted in the simulation of turbulent boundary layer deflection significantly due to the top-hat assumption. The shortcomings
[35], wind farms [36–37], and other fundamental problems. In the of the Jiménez model were also observed by Bastankhah et al. [11],
simulation, the atmospheric thermal stratification and the Coriolis force Howland et al. [17], Qian et al. [33], etc. For the BP model, the
are not considered. The actuator disk model [38–40] is used to represent streamwise and spanwise wake decay coefficients were set as ky = kz =
the wind turbine. For more details of numerical method, one can refer to 0.022 following Ref. [11]. Compared with the LES results, the BP model
the Appendix A. underestimates the wake deflection and this result is also supported by
The thrust coefficient perpendicular to the wind rotor and the ground Shapiro et al. [29]. Although the result of the QI model agrees well with
roughness are taken as CT = 0.75 and z0 = 0.1 m, respectively. The hub the LES result, it slightly underestimates the deflection of the wake in the
height (zh) and the rotor diameter (D) are assumed identical and are far wake region. In contrast, the Shapiro model outperforms the other
selected as zh = D = 100 m. The computational domain is 32D × 10D × three models according to the high-fidelity LES result. Hence, the Sha­
10D (Lx × Ly × Lz), in accordance with 256 × 128 × 120 (Nx × Ny × Nz) piro model is selected for the evaluation of the wind farm’s total power
grid points in streamwise, spanwise and vertical directions, respectively. output under yawed conditions. Since the rotational effect of the rotor
The turbine is placed 5D downstream of the inlet as shown in Fig. 1. A only have a slight influence on the lateral deflection of the turbine wakes
fully developed half channel flow is firstly generated under a constant under neutral atmospheric condition [42–43], the asymmetry of the
pressure gradient and then used as the inflow of the wind turbine. A wake trajectory between the positive and negative yaw angles is not
fringe region with a length of Lx/8 is set at the end of the computational considered.
domain to ensure that the flow can transit to the given inflow smoothly. According to Shapiro et al. [29], for a yawed turbine, the mean ve­
For details, one can refer to Ref. [41]. locity deficit in the wake can be computed as:
The non-dimensional profiles of the inflow velocity (U/Uh ) and ( )
√̅̅̅̅̅̅ D2 (y − yc (x) )2
the streamwise turbulence intensity (TIu = u’2 /Uh ) are shown in ΔU(x, y) = δu(x) 2 exp − (2)
8σ 0 2σ 2 (x)
Fig. 2, where Uh is the inflow velocity at the hub height. As expected,
the mean wind shear is stronger near the surface, leading to higher
where δu(x) is the mean velocity deficit in the streamwise direction at
streamwise turbulence intensity. The streamwise turbulence in­
the hub height, σ 0 = 0.235D is the proportional constant, σ (x) is the
tensity at the hub height is about 10%. Furthermore, Fig. 3 shows the
Gaussian width (standard deviation) of the wake, and yc (x) is the
normalized spectral density of the streamwise velocity (Euu D− 1 u−* 2 ) at deflection of the wake center. The term δu(x) can be calculated as:
the hub height as a function of kx D, where kx is the streamwise [ (√̅̅̅ ) ]
wavenumber. It can be observed that the energy spectrum follows the δu0 1 2x
δu(x) = 2 1 + erf (3)
− 5/3 law for the inertial subrange and the − 1 law for the production dw (x) 2 D
subrange, which indicates that the in-house code can produce real­
istic turbulent inflow. and
Fig. 4 shows the contours of the streamwise velocity for γ = 30◦ , { ( √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ )
where γ is the yaw angle of the turbine. As can be observed, for the δu0 = U0 1 − 1 − CT cos2 γ (4)
yawed turbine, the wake exhibits an obvious deflection and a significant dw (x) = 1 + kw ln(1 + exp[(x − 2R)/R ] )
velocity deficit is induced due to the shelter effect of the turbine.
Following Refs. [17,27], the center of the wake can be computed as: where δu0 is the initial velocity deficit in the streamwise direction, dw (x)
is the width of turbine wake, CT is the thrust coefficient of the rotor, γ is
the yaw angle and kw is the wake decay coefficient. In Eq. (2), σ(x) =

Fig. 1. Schematic diagram of standalone wind turbine with a yaw angle of γ.

3
H. Ma et al. Applied Energy 303 (2021) 117691

Fig. 2. The profiles of (a) the non-dimensional inflow velocity and (b) the streamwise turbulence intensity.

σ0 dw (x), and the term yc (x) can be computed as:


⎧ ∫ x ′

⎪ − δv(x ) ′

⎪ yc (x) = dx

⎪ − ∞ U0



⎨ [ (√̅̅̅ ) ]
δv(x) =
δv0 1
1 + erf
2x (5)

⎪ dw2 (x) 2 D






⎩ 1
δv0 = U0 CT cos2 γsinγ
4

where δv(x) is the mean velocity deficit in the spanwise direction at the
hub height and δv0 is the initial velocity deficit in the spanwise direction.

2.2. Superposition of the wakes and the power output

In the alignment direction, the turbines strongly interact with each


other through their wakes, and the mean velocity of the flow can be
calculated by a superposition method based on a single wake model.
Currently, two main types of superposition approaches, i.e., the “sum of
the velocity deficits” [44–45] and the “sum of squares of the velocity
deficits” [46–47], are available. In this study, the most widely used
“square sum” superposition method is adopted and for the i-th wind
turbine:
Fig. 3. Normalized streamwise spectra of streamwise velocity versus nondi­
mensional wavenumbers at the hub height for the inflow.

Fig. 4. Contour of non-dimensional mean velocity on the x-y plane at the hub height (white dotted lines denote the boundary of the stream tube through the wind
rotor, black dotted line represents the wake centerline).

4
H. Ma et al. Applied Energy 303 (2021) 117691

Fig. 5. Comparison of deflection of the wake center with the wake models and LES.

√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
√ i− 1 ( mutation in the process of natural selection and natural inheritance. The
Ui √∑ ΔUki )2
=1-√ (6) algorithm is usually used to address the multi-variable optimization
U0 k=1
U0 problems and exhibits excellent global optimization capabilities with
inherent implicit parallelism. The algorithm can adaptively adjust the
where U0 denotes the incoming velocity, Ui represents the mean wind search direction with high efficiency. The five issues, i.e., the parameter
speed of the i-th turbine, and ΔUki is the mean velocity deficit on the i-th coding, the initial population setting, the fitness function design, the
turbine due to the shelter effect of the k-th turbine, which can be genetic operation design, and the control parameter setting, constitute
calculated by Eq. (2). the key content of the optimization procedure. To speed up the
For a non-waked and non-yawed turbine, the power output is: convergence and obtain accurate results rather than the local optimum,
1 the GA with elite selection is adopted in the study. The parameters used
P0 = ρACp U03 (7) are shown in Table 1.
2
The optimization procedure is as follows:
For the i-th turbine with a yaw angle of γi , the power is computed as
[16,34]: Step 1: Generate the initial yaw angle aggregation groups randomly.
1 In each group, there are Ps yaw angle aggregation individuals. The
Pi = ρACp Ui3 cos3 γ i (8) range of the yaw angle is − 40◦ to 40◦ with a minimal interval of 0.3◦ .
2
Step 2: Encode the yaw angle aggregate group individuals using the
where ρ is the density, A is the swept area of the wind rotor, and Cp is the binary method. The encoding bit is 8N-8.
power coefficient. The normalized power output of the i-th turbine can Step 3: Calculate the value of the objective function and the fitness
be denoted as: function of each aggregate group individual.
( )3 Step 4: Duplicate and cross-operate the parent group: select good
Pi 12 ρACp Ui3 cos3 γi Ui
Pi0 = = 1 3
= cosγi (9) individuals to retain to the next generation through the roulette
P0 2
ρACp U0 U0
method, and exchange the two individuals with a probability of Pc
And the normalized total power can be computed as: from a specific position in their respective codes.
Step 5: Perform mutation operation to the parent population: each
∑ bit of the code of every individual in the parent is reversed with a
N
Ptot = Pi0 (10)
i=1 probability of Pm.
Step 6: Use the individual with the maximum fitness in the parent to
where N is the total number of the aligned turbines. replace the individual with the minimum fitness in the offspring
population to ensure that the direction of evolution does not reverse.
2.3. The optimization algorithm and procedure Step 7: Determine whether the optimization converges: when the
number of individuals with the maximum fitness exceeds 95% of the
For the cooperative yaw control, the maximization of total individuals, the process converges and the optimal individual
Ptot =
∑N results can be outputted as the optimized result; otherwise, return to
i=1 Pi0 is taken as the optimization goal, and the yaw angles
γ = (γ1 , γ 2 , ⋯, γN− 1 , γN ) is taken as the optimization variables. The step 2 and continue the iteration.
fitness function VF is selected as:
3. Cooperative yaw control strategy

N ∑
N
VF = Ptot y − Ptot = Pi0 |γi =γi − Pi0 |γi =0 (11)
For aligned wind turbines, many factors such as the number of tur­
ny
i=1 i=1
bines (N), the turbine spacing (Sx ), the thrust coefficient (CT ), and the
where Ptot y and Ptot ny are the normalized total power with and without wake decay coefficient (kw ) can significantly influence the cooperative
cooperative yaw control. The widely used Genetic algorithm (GA) is yaw control strategy. In this section, the influence of the above factors
adopted for the optimization. on the cooperative yaw control strategy is systematically analyzed, and a
GA is a search algorithm based on the natural selection and popu­ more generalized and efficient method to determine the yaw angles of
lation genetic mechanism. It simulates reproduction, hybridization, and the turbines for the wind farm control is proposed.

5
H. Ma et al. Applied Energy 303 (2021) 117691

Table1
Parameters for the Genetic algorithm.
N Population length Crossover probability Pc Mutation probability Pm Genetic Algebra Generation

N =3 40 0.70 0.08 150


N =4 40 0.70 0.08 200
N =5 50 0.75 0.10 200
N =6 50 0.75 0.10 250

3.1. Simplification of the optimization space except the last one which does not need to yaw. In other words, the wake
losses of the aligned turbines can be more pronouncedly reduced when
To simplify the physical problem, the influence of N on the optimized the wakes of the upstream turbines deflect to the same direction; (iii)
yaw angles is firstly examined through a series of combinations of the The difference of the yaw angles between the second and (N-1)-th tur­
thrust coefficients, wake decay coefficients, and turbine spacings. bine can be negligible, and the yaw angles decrease with wake decay
Following the previous studies [48–50], we set the thrust coefficient CT coefficient overall. (iv) Compared with the second to (N-1)-th turbines,
= 0.75 and the turbine spacing Sx = 7D. The yaw angles are optimized the yaw angle of the first turbine is smaller.
for different turbine numbers (N = 3–6) and different wake decay co­ Fig. 6(d) shows the optimized yaw angles for kw = 0.075. For these
efficients (kw = 0.03, 0.05 and 0.075). Fig. 6(a) shows the variation of cases with larger kw, the total wake effect is much weaker and the first
the control efficiency η with the wake decay coefficient, where turbine does not need to yaw. The optimized yaw angles of the turbines
η=ΔP/P × 100%, in which ΔP denotes the added power by the coop­ become not sensitive to N until N > 5. In view of the slight control ef­
erative control and P is the total power of the aligned turbines without ficiency for kw = 0.075, the cases with small wake decay coefficients
yaw control. As can be observed, the total power of the turbines is (kw =0.03 to kw = 0.05) that are usually recommended for the offshore
enhanced significantly due to the cooperative control. And the control wind farms [51] are mainly focused in the present study.
efficiency increases with N and decreases with the wake decay coeffi­ As described above, it can be assumed that for a certain kw and Sx , the
cient. For N = 6 and kw = 0.03, the efficiency can achieve 46% while for yaw angles of the optimized cooperative control has the certain form of
N⩽5 and kw = 0.075, the efficiency is only less than 4%. γ = (γ1 , γ2 , ……, γ 2 , 0), where γ1 is the yaw angle of the first turbine, and
Fig. 6(b) and (c) show the optimized yaw angles for kw = 0.03 and kw γ 2 is the yaw angle of the second turbine to the (N-1)-th turbine. Table 2
= 0.05, respectively. For a specific kw, four different N (N = 3, 4, 5, 6) are compares the optimized results from the original strategy and the
considered. From the figures, we can summarize the characteristics of simplified strategy. As is shown, the control efficiency of the simplified
the control strategy as follows. (i) For a certain wake decay coefficient, strategy is very close to that of the original control strategy. The effi­
the yaw angles of the aligned turbines are not sensitive to N. (ii) In the ciency degradation induced by the simplification is negligible. In
optimized control, all the turbines should yaw to the identical direction particular, when N = 6, the deviation is less than 0.1%. This

Fig. 6. Control efficiency and optimized yaw angles for different turbine numbers. (a) Variation of the control efficiency with kw for CT = 0.75 and Sx = 7D, (b)
optimized yaw angles for different turbine numbers with kw = 0.03, (c) optimized yaw angles for different turbine numbers with kw = 0.05, and (d) optimized yaw
angles for different turbine numbers with kw = 0.075.

6
H. Ma et al. Applied Energy 303 (2021) 117691

Table 2
Comparison of original optimized control strategy and the simplified strategy.
η0 is the control efficiency of the original strategy, η1 is the control efficiency of
the simplified strategy, Δη = η0 − η1 , only the optimized yaw angles for the
simplified strategy are shown in the Table.
kw Sx/D N γ1 γ2 η0 (%) η1(%) Δη (%)

0.03 7 3 23.0 31.6 11.2 8.9 2.3


4 23.4 19.8 3.6
5 35.0 30.7 4.3
6 46.2 46.2 0.0

0.04 7 3 19.5 30.4 6.3 6.2 0.1


4 13.8 13.8 0.0
5 20.6 20.1 0.5
6 26.7 26.7 0.0

0.05 7 3 13.8 28 3.5 1.5 2.0


4 8.0 6.2 1.8
5 12.0 10.2 1.8
6 15.4 15.4 0.0

Fig. 7. Variation of the optimized control efficiency with CT. Black, blue and
demonstrates that the simplification is reasonable and the optimization red lines represent kw = 0.03, 0.04 and 0.05, respectively, and the solid, dashed
variables can be greatly reduced without deviating much from the and dotted lines represent Sx = 6D, 7D and 8D, respectively. (For interpretation
optimal results. Therefore, only the two yaw angles γ 1 and γ2 are of the references to colour in this figure legend, the reader is referred to the web
required to be determined for the strategy of the cooperative control. version of this article.)

has an obvious inflection point and the slope of the curve decreases with
3.2. Generalized formulas to determine the yaw angles CT . Fig. 8(b) plots the variation of γ1 with kw and a clear negative cor­
relation can be observed between the two variables. This implies that
To reveal the dependence of the optimized yaw angles on the thrust with the increase of the wake decay coefficient, the shelter effect of the
coefficient, the wake decay coefficient and the turbine spacing, four first turbine is reduced, and at this condition, only a smaller yaw angle is
aligned turbines are considered. The control strategies are optimized for appropriate to regulate the wakes. Fig. 8(c) further illustrates the vari­
27 combinations of the above parameters (CT =0.65, 0.75 and 0.85, ation of γ 1 with the turbine spacing. No significant effect of the turbine
kw =0.03, 0.04 and 0.05, Sx =6D, 7D and 8D). For more practical rele­ spacing on γ1 can be observed, which indicates that γ1 is independent on
vance, all the parameters are selected based on actual wind farms the arrangement of downstream turbines.
[16,25,45,51] and the combinations of the parameters are shown in To obtain a simple formula for γ1 , the combined variable CT /kw is
Table 3. Fig. 7 shows the variation of the control efficiency with thrust introduced. Fig. 9 shows the variation of γ 1 with CT /kw . The strategies of
coefficient for different turbine spacings and wake decay coefficients. the cooperative control with the efficiency less than 5% are marked in
Similar results are obtained as Section 3.1, the stronger the wake effect blue while the others are marked in black. As can be observed, the three
(with larger thrust coefficient, smaller turbine spacing, and smaller blue points in the lower left corner have a severe deviation with the
wake decay coefficient), the higher efficiency of the cooperative control. other points. It is noted that the less of CT /kw , the weaker of the wake
For most of the cases, the control efficiency exceeds 10% except for cases effect, and a lower control efficiency will be obtained through the
that the wake losses are relatively small. It is also should be mentioned cooperative control. For the three points, we can only get an efficiency
that the yaw control efficiency is more sensitive to the wake decay co­ less than 3% which exactly corresponds to the lower left points in Fig. 7.
efficient than the thrust coefficient and turbine spacing. Therefore, Except the three points, γ1 shows an approximate linear relation with
applying the optimization strategy on wind farms with a small wake CT /kw . Hence, the severely deviated data (the data in the lower left
decay coefficient, such as offshore wind farms with low ambient tur­ corner) are ignored and the scatters can be approximately fitted linearly
bulence intensity, is expected to achieve large control efficiency. as:
Fig. 8(a) shows the variation of the yaw angles of the first turbine γ 1
with CT for the optimized control strategy. As is shown, γ1 monotonously γ 1 = k1
CT
+ c1 (12)
increases with the thrust coefficient. This indicates that, for larger CT, kw
the velocity deficit of the first turbine is greater, so a larger yaw angle is
where k1 = 0.872 and c1 = 0.952. The optimized yaw angle of the first
needed to reduce the shelter effect. For kw = 0.03 and 0.04, the γ1 − CT
turbine for any CT and kw can be quickly calculated by Eq. (12).
curves are approximately linear, while for kw = 0.05, the γ1 − CT curve
Although it is not a perfect fitting, the error introduced by the linear
⃒ ⃒
⃒ ⃒
Table 3 fitting is engineering acceptable. We define Δγ = ⃒γ mod − γfit ⃒, where
Combinations of the thrust coefficients (CT), the wake decay coefficients (kw) the γ mod and γfit represent the scatters and the fitting value, respectively.
and the turbine spacings (Sx) for optimization of the cooperative yaw control The largest and the smallest Δγ are 3.12◦ (at CT /kw = 21.25) and 0.25◦
strategy. (at CT /kw = 25.00), respectively.
Case CT kw Sx/D Fig. 10(a)-(c) show the variation of γ 2 with CT , kw and Sx for the
1–3 0.65 0.03 6–8 optimized control, respectively. Similar to γ 1 , γ 2 increases with CT , and
4–6 0.65 0.04 6–8 decreases with kw and Sx . For cases with small kw , large CT and small Sx ,
7–9 0.65 0.05 6–8 large yaw angles are needed to reduce the wake losses. This indicates
10–12 0.75 0.03 6–8
that when the shelter effect of the second turbine to the (N-1)-th turbine
13–15 0.75 0.04 6–8
16–18 0.75 0.05 6–8
is strong, large yaw angles should be set to obtain significant wake de­
19–21 0.85 0.03 6–8 flections for more power production.
22–24 0.85 0.04 6–8 Fig. 11 shows the variation of γ2 with the combined variable
25–27 0.85 0.05 6–8 CT /(kw Sx ). A strong positive correlation exists between the two

7
H. Ma et al. Applied Energy 303 (2021) 117691

Fig. 8. (a) Variation of γ1 with CT , black, blue and red lines represent kw = 0.03, 0.04 and 0.05, respectively, and the solid, dashed and dotted lines represent Sx = 6D,
7D and 8D, respectively; (b) variation of γ1 with kw , black, blue and red lines represent Sx = 6D, 7D and 8D, respectively, and the solid, dashed and dotted lines
represent CT = 0.65, 0.75 and 0.85, respectively; (c) variation of γ1 with Sx , black, blue and red lines represent kw = 0.03, 0.04 and 0.05, respectively, and the solid,
dashed and dotted lines represent CT = 0.65, 0.75 and 0.85, respectively. (For interpretation of the references to colour in this figure legend, the reader is referred to
the web version of this article.)

By now, the generalized cooperative control strategy for aligned


turbines is proposed for different CT, kw and Sx. For a realistic wind farm,
the parameters CT and Sx can be obtained from the thrust curve of the
turbines and the layout of the wind farm. The only parameter to be
determined is the wake decay coefficient kw which can be approximately
determined using the SCADA data of the wind turbines. In particular, the
mean power and wind speed of the second turbine can be obtained from
the operation data, and then the equivalent kw can be calculated using
Eqs. (2)-(5) under the fully waked condtion.

Fig. 9. Variation of γ1 with CT /kw .

variables. Although the data points are relatively scattered, the points
can be approximately fitted as a linear function in engineering:
CT
γ 2 = k2 + c2 (13)
kw Sx

where k2 = 4.697 and c2 = 16.64. Therefore, Eq. (13) can be used to


compute the yaw angles of the second turbine to the (N-1)-th turbine for
the optimized control. As is shown in Fig. 11, the largest and the smallest
Δγ between the fitting line and the scatters are 6.25◦ (at CT /(kw Sx ) =
Fig. 11. Variation of γ2 with CT /(kw Sx ).
2.833) and 0.61◦ (at CT /(kw Sx ) = 4.167), respectively.

Fig. 10. (a) Variation of γ2 with CT , (b) variation of γ2 with kw and (c) variation of γ2 with Sx . The line-colors and line-types are identical to that in Fig. 8.

8
H. Ma et al. Applied Energy 303 (2021) 117691

4. Validation and discussion turbines’ power can gain much more power revenue from the down­
stream turbines by the cooperative control. Especially for the last tur­
To validate the proposed formulas, high-fidelity LES studies are bine, a conspicuous power revenue can be obtained. Due to the yaw
performed. In the simulation, five turbines are placed in the streamwise angles of the upstream turbines, the trajectories of the turbine wakes are
direction above a flat wall. Two thrust coefficients (CT =3/4 and 8/9) deflected, which significantly reduces the shelter effect of the upstream
and two turbine spacings (Sx =6D and 7D) are considered. The ground turbines. Note that the yaw angle of the last turbine is kept zero, hence,
roughness z0 of the aligned turbines is 0.1 m. The rotor diameter D is the turbine can gain power revenue due to the weakened wakes but
assumed identical to the hub height zh, and we set D = zh = 100 m. The without power loss induced by the yaw of itself.
parameters for all the validation cases are summarized in Table 4. The Fig. 14 demonstrates the contours of the mean velocity at the hub
symbol “A” is used for the type of cases with Sx = 7D, and CT = 3/4. height for the six validation cases. The deflections of the turbine wakes
While for the type B cases, the turbine spacing is kept identical to that of can be obviously observed due to the cooperative yaw control. For the
the type A cases but the thrust coefficient is changed to 8/9. For type C second to the fourth turbine, although their yaw angles are set identical,
cases, the thrust coefficient is set identical to that of the type A cases but the deflection of their wakes gradually increases due to the increasing
the turbine spacing is changed to 6D. The number ‘0’ is used to label the asymmetry of the incoming flow. To quantify the deflection of the tur­
reference cases, i.e., the aligned wind turbines without the cooperative bine wakes, the deflection at the position of the downstream adjacent
control, and ‘1’ is used for the cases using the proposed control strategy. turbine is calculated and shown in Table 5. The results indicate that from
The inhouse code LESGO is used for the simulations. The computational the third turbine, the offset between the rotor center and the trajector­
domain is 50D × 10D × 10D (Lx × Ly × Lz), in accordance with 256 × y of the upstream turbine’s wake center reaches or exceeds 0.5D, which
128 × 120 (Nx × Ny × Nz) grids in the streamwise, spanwise and vertical implies that the downstream turbine effectively avoids the centerline of
directions, respectively. A fully developed turbulence flow with TIu = the upstream turbine wake through the cooperative yaw control. The
10% (corresponds to the wake decay coefficient kw = 0.05 [29]) is contours also show that the main wake region is limited within the 2D
generated by a precursor method as the inflow of the aligned turbines, range of the turbine in the spanwise direction. Therefore, the control
where TIu is the streamwise turbulence intensity at the hub height. strategy proposed for aligned wind turbines in this study is still appli­
Correspondingly, the wake decay coefficient is set as 0.05 to compute cable to turbine arrays with spanwise spacing greater than 4D. Thus, the
the yaw angles for the cooperative control. formulas can be used when the wind direction is parallel to the align­
As shown in Table 4, the LES results further demonstrate that the ment direction of the turbines. It should be mentioned that aligned
total power of the aligned turbines can be significantly enhanced by the turbines are often can be observed in many realistic wind farms, e.g., the
proposed control strategy. But it is worth noting that the control effi­ Horns Rev wind farm in Denmark, the Dafeng offshore wind farm in
ciency obtained from LES is remarkably less than the efficiency pre­ China, etc., hence the proposed strategy is expected to have a wide range
dicted by the wake model. This overestimation is more pronounced for of application scenarios.
larger CT, such that for the type B cases, the control efficiency from the Fig. 15 demonstrates the contours of the instantaneous velocity at
wake model is about 33.3%, while the efficiency from LES is only 17.5%. hub height for Case B0 and Case B1. For brevity, the instantaneous flow
This might because the variation of the wake decay coefficient of field of the other cases are not plotted since they show similar charac­
downstream turbines due to the turbulence intensity variation is teristics as the representative cases. As is shown in Fig. 15, due to the
neglected and the wake losses are mispredicted by using only a constant unsteady characteristics of the ABL and its interaction with the wind
wake decay coefficient in the assessment of the wake effect. To illustrate turbines, the wakes show a clear meandering phenomenon. Under yaw
this further, the contour of the turbulence intensity for Case B0 and B1 conditon, the large scale wake structures exhibit a clear directional
are plotted in Fig. 12. As is shown, due to the superposition of the tur­ deflection, which agrees well with Fig. 14. The results implies that the
bine wakes, the turbulence intensity shows an obvious augment on the yaw angles shift the mean center line of the meandering, and then
downstream turbine positions for the non-yawed case, and this variation reduce the wake effect of the alligned turbines.
on the turbulence intensity can significantly accelerate the recovery of It should be noted that, apart from the overall power increment of a
the turbine wakes [34,37]. While for Case B1, the turbulence intensity wind farm, the intelligent yaw control can remarkably influence the
even shows a slight decrease for the downstream turbines except the last turbine fatigue loads in practical applications. Fleming et al. [18] found
one. Therefore, the adoption of a constant wake decay coefficient could that, the blade edgewise moments which are mainly dominated by
introduce a substantial error here. Even so, the effectiveness of the gravity forces are not heavily coupled with yaw-misalignment. How­
proposed formulas is well validated. ever, due to the deflection of the upstream tubine wakes under yaw-
Fig. 13 shows the variation of the power of the turbine with the serial misalignment, the downstream turbine will experience unsteady asym­
number of the turbines from upstream to downstream. Both the cases metric loading [52–54], which can significantly increase the flapwise
with and without the cooperative control are plotted for comparison. As bending moments, the drivetrain torsion, and the tower-base bending
can be observed, due to the yaw control, the power of the turbines ex­ moments of the downstream turbines [18]. Furthermore, Kanev and
hibits a monotonous rise with downstream from the second one. The Savenije [54] announced that the Damage Equivalent Loads (DEL) can
powers of the upstream turbines (the first turbine and the second one) be enhanced by 15–20% under the partial wake overlap condition when
are slightly reduced, while the power of the downstream ones are sub­ the wake center passing through the edge of the rotor. Therefore, the
stantially enhanced. In other words, the slight sacrifice of the upstream influence of the cooperative yaw control on the structural loads is very
worthy to be investigated in the future research.
Table 4
Numerical settings and results of the validation cases. ηLES denotes the control 5. Summary and conclusions
efficiency from LES, and ηmod denotes the control efficiency predicted by the
wake model. In this study, the cooperative yaw control of aligned turbines for
Case CT Z0/m Sx/D ηLES ηmod power maximization is systematically studied under neutral atmo­
spheric condition. Two generalized and effective algebra formulas to
A0 3/4 0.1 7 – –
A1 3/4 0.1 7 9.3% 11.9%
determine the turbines’ yaw angles are proposed for the cooperative
B0 8/9 0.1 7 – – control of a wind farm. The strategy is then validated by high-fidelity
B1 8/9 0.1 7 17.5% 33.3% large-eddy simulation (LES). The main conclusions are as follows:
C0 3/4 0.1 6 – –
C1 3/4 0.1 6 9.9% 14.0%

9
H. Ma et al. Applied Energy 303 (2021) 117691

Fig. 12. Contours of turbulence intensity at the hub height for (a) Case B0, and (b) Case B1.

Fig. 13. Variation of the normalized wind turbine power from upstream to downstream. (a) Case A0 and A1, (b) Case B0 and B1, (c) Case C0 and C1.

(1) Cooperative yaw control can substantially increase the total and all the turbines should yaw to the identical direction except
power of the aligned turbines. Higher control efficiency can be the last turbine. The cooperative yaw control strategy can be
obtained for cases with stronger wake effect (larger thrust coef­ simplified to a solution space only containing two yaw angles γ =
ficient, smaller turbine spacing and smaller wake decay coeffi­ (γ1 , γ2 ……γ2 , 0), where γ1 is the yaw angle of the first turbine and
cient). For the wake decay coefficient kw = 0.03–0.05, the γ2 is the yaw angle of the second turbine to the (N-1)-th turbine.
optimized yaw angles are not sensitive to the turbine number N,

10
H. Ma et al. Applied Energy 303 (2021) 117691

Fig. 15. Contours of instatanious streamwise velocity u/Uh on the horizontal


plane at the hub height for (a) Case B0, (b) Case B1. Uh represents inflow ve­
locity at the hub height.

(2) For the optimized control strategy, γ1 increases with the thrust
coefficient, and decreases with the wake decay coefficient. γ1
exhibits a strong positive correlation with the combined variable
CT /kw and can be approximately computed as γ1 = k1 CT /kw + c1 ,
where k1 = 0.872 and c1 = 0.952. While γ2 can be approximately
solved as γ 2 = k2 CT /(kw Sx ) + c2 , where k2 = 4.697 and c2 =
16.64.
(3) To validate the proposed formulas, high-fidelity LESs are per­
formed for five aligned turbines with different thrust coeffects
and turbine spacings over a flat ground. The results show that the
yaw angles enable the downstream turbines to effectively avoid
the wake trajectory center of the upstream turbines, which
significantly increased the total power of the wind farm. A
notable control efficiency of 17.48% can be obtained for the
aligned turbines with CT = 8/9 and Sx = 7D.

Some aspects of the cooperative yaw control are still worthy to be


further investigated, such as the change of the fatigue loads of the tur­
bines, the rotor rotating effects, the influence of the Coriolis force, etc. In
the following work, we will focus on the cooperative control of a wind
farm subjecting to various wind conditions, and demonstrate the control
Fig. 14. Contours of mean streamwise velocity U/Uh on the horizontal plane at strategy on realistic offshore wind farms.
the hub height for (a) Case A0, (b) Case A1, (c) Case B0, (d) Case B1, (e) Case
C0, and (f) Case C1. Uh represents inflow velocity at the hub height. CRediT authorship contribution statement

Hongliang Ma: Data curation, Writing – original draft. Mingwei Ge:


Table 5
Deflection of the wake center at the position of the downstream adjacent Conceptualization, Methodology, Software, Supervision. Guangxing
turbine. Wu: Writing - review & editing. Bowen Du: Writing - review & editing.
Yongqian Liu: Conceptualization.
Case Turbine serial No. Wake deflection

A1 1 0.30 Declaration of Competing Interest


2 0.52
3 0.75
4 0.80 The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
B1 1 0.40
2 0.60
the work reported in this paper.
3 0.80
4 1.00 Acknowledgements
C1 1 0.25
2 0.50 The research is supported by National Key R&D Program of China
3 0.60 (No. 2019YFE0104800), the National Natural Science Foundation of
4 0.72 China (No. 11772128) and the State Key Laboratory for Alternative
Electrical Power System with Renewable Energy Sources (No.
LAPS202107).

11
H. Ma et al. Applied Energy 303 (2021) 117691

Appendix A. Numerical method

A-1. Large-eddy simulation methodology

In the present simulation, the atmospheric thermal stratification is not considered, and hence the continuity equation and the filtered incom­
pressible Navier-Stokes (N-S) equations for neutral flow condition are taken as the governing equations for LES:
∂̃ui
=0 (A1)
∂xi

∂̃ui ∂̃ui p* ∂τij fi


1 ∂̃ ∂1 p∞
uj
+ ̃ = − − − +δ (A2)
∂t ∂xj ρ ∂xi ∂xj ρ i1 ρ

* *
where, ̃ui represents filtered velocity field, t is time; ρ is the density of the fluid; ̃ p is the filtered modified pressure, which equals to ̃ p = ̃p+
d
ρτkk /3 - p∞ /ρ; τij represents sub-grid stress, τij = ũ
i uj − ̃
ui ̃
uj , and τij = τij − δij τkk /3 is its deviatoric part which is solved by a Lagrangian scale dependent
dynamic Smagorinsky model; fi is the body force which is used to model the effect of the wind turbine on the flow; δij is the Kronecker delta function;
∂1 p∞ /ρ denotes the pressure gradient in the streamwise direction, and we set ∂1 p∞ /ρ = u2* /H to generate the pressure-driven flow perpendicular to the
wind rotor. In the atmospheric boundary layer, the Reynolds number is very high and the wind turbine effects are parametrized, hence the molecular
dissipation and viscous stress are neglected following many previous researches [55–57].
In the numerical method, the vertically staggered grids are adopted. For spatial discretization, the pseudo spectral method is used and thus periodic
boundary conditions are used in the horizontal directions. The second-order central finite difference method is used in the vertical direction. For time
advancement, the second order Adams-Bashforth method is adopted. the stress-free boundary condition is adopted for upper boundary and the
classical wall stress boundary condition is used for the bottom wall, in which the instantaneous wall stress is connected to the velocity of the first grid
point through the standard logarithmic law, and calculated by the following equation:
( )2 ( )
κ
(A3)
2 2
τw (x, y) = − u +̃
̃ v
ln[z/z0 ]

where u* is the friction velocity; ̃


ur is the local filtered horizontal velocity of the first horizontal plane (z = Δz/2); ̃
u and ̃
v refer to the local average
velocities filtered by the scale of 2Δ.
̃ z0 represents the aerodynamic ground roughness; κ = 0.4 is the von Kármán constant.

A-2. Actuator disk model

The actuator disk model is adopted to model the wind turbine, in which the rotation effect of wake is neglected and only the thrust force generated
by the wind turbine is considered. It is assumed that the axial thrust is evenly distributed on the disk. Although the flow structure of near wake cannot
be obtained using this model, it causes very small error in the far wake region. The total thrust force acting on the wind turbine is:
1 2 π 2 1 Ud2 π 2
Ft = − ρC T U ∞ D = − ρCT D (A4)
2 4 2 (1 − a)2 4

where, CT is the thrust coefficient, U∞ is the velocity of the incoming flow, Ud = U∞ (1 − a) is the wind speed at the disk, in which a is the axial
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅
induction factor and a = (1 − 1 − CT )/2.
The wind speed Ud on the rotor is taken as the reference velocity for the wind rotor’s thrust force [58]. Further, an average disk velocity is used to
modeling the thrust forces acting on the fluid due to its interaction with the rotating blades. The flow velocity on the rotor can be obtained by
〈 〉
averaging the flow field of the rotor plane in a certain time period T, i.e., uT . The total thrust on the rotor is computed as:
d

1 ′ 〈 〉2 π
Ft = − ρCT uT d D2 (A5)
2 4

where CT is the modified thrust coefficient and is normally set as CT = 1.33, in accordance with CT = 0.75 and a = 0.25. While for Case B0 and B1, the
′ ′

thrust coefficient is set as CT = 2 corresponding to CT = 8/9 and a = 1/3.


References [4] Wilson DG, Rodrigues S, Segura C, Loshchilov I, Hutter F, Buenfil GL, et al.
Evolutionary computation for wind farm layout optimization. Renewable Energy
2018:681–91.
[1] Archer CL, Vasel-Be-Hagh A, Yan C, Wu S, Pan Y, Brodie JF. Review and evaluation
[5] Mahulja S, Larsen GC, Elham A. Engineering an optimal wind farm using surrogate
of wake loss models for wind energy applications. Appl Energy 2018:1187–207.
models. Wind Energy 2018;21(12):1296–308.
[2] Thomsen K, Madsen HA, Larsen GC, Larsen TJ. Comparison of methods for load
[6] Bansal JC, Farswan P. Wind farm layout using biogeography based optimization.
simulation for wind turbines operating in wake. J Phys Conference 2007;75:
Renewable Energy 2017:386–402.
012072.
[7] Cruz LEB, Carmo BS. Wind farm layout optimization based on CFD simulations.
[3] Hou P, Hu W, Soltani M, Chen C, Chen Z. Combined Optimization for Offshore
J Braz Soc Mech Sci Eng 2020;42(8).
Wind Turbine Micro Siting. Appl Energy 2017:271–82.
[8] Beck CJ, Romero ACH. Wind farm layout optimization on complex terrains -
Integrating a CFD wake model with mixed-integer programming. Appl Energy
2016.

12
H. Ma et al. Applied Energy 303 (2021) 117691

[9] Bossi NK, Porté-agel F. Realistic Wind Farm Layout Optimization through Genetic [33] Frandsen S, Barthelmie R, Pryor S, Rathmann O, Larsen S, Højstrup J, et al.
Algorithms Using a Gaussian Wake Model. Energies 2018;11(12). Analytical modelling of wind speed deficit in large offshore wind farms. Wind
[10] Stanley AP, Ning A, Dykes K. Optimization of Turbine Design in Wind Farms with Energy 2016;9(1–2):39–53.
Multiple Hub Heights, Using Exact Analytic Gradients and Structural Constraints. [34] Guo-Wei Q, Takeshi I. A New Analytical Wake Model for Yawed Wind Turbines.
Wind Energy 2019;22(5):605–19. Energies 2018;11(3):665.
[11] Bastankhah M, Porté-agel F. Experimental and theoretical study of wind turbine [35] Yang XIA, Sadique J, Mittal R, Meneveau C. Integral wall model for large eddy
wakes in yawed conditions. J Fluid Mech 2016:506–41. simulations of wall-bounded turbulent flows. Phys Fluids 2015;27(2):025112.
[12] Marathe N, Swift A, Hirth BD, Walker R, Schroeder J. Characterizing power [36] Calaf M, Meneveau C, Meyers J. Large eddy simulation study of fully developed
performance and wake of a wind turbine under yaw and blade pitch. Wind Energy wind-turbine array boundary layers. Phys Fluids 2010;22(1):015110.
2016;19(5):963–78. [37] Ge M, Yang H, Zhang H, et al. A prediction model for vertical turbulence
[13] Annoni J, Gebraad PM, Scholbrock AK, Fleming PA, Van Wingerden JW. Analysis momentum flux above infinite wind farms. Phys Fluids 2021;33(5):055108.
of axial-induction-based wind plant control using an engineering and a high-order [38] Ge M, Wu Y, Liu Y, et al. A two-dimensional model based on the expansion of
wind plant model. Wind Energy 2016;19(6):1135–50. physical wake boundary for wind-turbine wakes. Appl Energy 2019;233:975–84.
[14] Grant I, Parkin P, Wang X. Optical vortex tracking studies of a horizontal axis wind [39] Ge M, Wu Y, Liu Y, et al. A two-dimensional Jensen model with a Gaussian shaped
turbine in yaw using laser-sheet, flow visualisation. Exp Fluids 1997;23(6):513–9. velocity deficit. Renewable Energy 2019;141:46–56.
[15] Grant I, Parkin P. A DPIV study of the trailing vortex elements from the blades of a [40] Bastankhah M, Li Z, Yang X. Evaluation of Actuator Disk Model Relative to
horizontal axis wind turbine in yaw. Exp Fluids 2000;28(4):368–76. Actuator Surface Model for Predicting Utility-Scale Wind Turbine Wakes. Energies
[16] Jiménez Á, Crespo A, Migoya E. Application of a LES technique to characterize the 2020;13.
wake deflection of a wind turbine in yaw. Wind Energy 2009;13(6):559–72. [41] Nordström J, Nordin N, Henningson D. The Fringe Region Technique and the
[17] Howland MF, Bossuyt J, Martínez-Tossas LA, Meyers J, Meneveau C. Wake Fourier Method Used in the Direct Numerical Simulation of Spatially Evolving
structure in actuator disk models of wind turbines in yaw under uniform inflow Viscous Flows. Soc Ind Appl Math 1999.
conditions. J Renewable Sustain Energy 2016;8(4). [42] Porté-Agel F, Lu H, Wu Y. A large-eddy simulation framework for wind energy
[18] Fleming P, Gebraad PM, Lee S, Van Wingerden JW, Johnson K, Churchfield M, applications. In Fifth International Symposium on Computational Wind
et al. Simulation comparison of wake mitigation control strategies for a two- Engineering, Chapel Hill; 2010.
turbine case. Wind Energy 2015;18(12):2135–43. [43] Lin M, Porté-Agel F. Large-Eddy Simulation of yawed wind-turbine wakes:
[19] Fleming P, Gebraad PM, Lee S, Van Wingerden JW, Johnson K, Churchfield M, Comparisons with wind tunnel measurements and analytical wake models.
et al. Evaluating techniques for redirecting turbine wakes using SOWFA. Energies 2019;12(23):4574.
Renewable Energy 2014:211–8. [44] Lissaman PBS. Energy effectiveness of arbitrary arrays of wind turbines. J Energy
[20] Adaramola MS, Krogstad P. Experimental investigation of wake effects on wind 1979;3.6:323–8.
turbine performance. Renewable Energy 2011;36(8):2078–86. [45] Amin N, Porté-Agel F. Analytical Modeling of Wind Farms: A New Approach for
[21] Bartl J, Franz Mühle, Lars Sætran. Wind tunnel study on power and loads Power Prediction. Energies 2016;9(9):741.
optimization of two yaw-controlled model wind turbines. 2018. [46] Katic I, Hojstrup J, Jensen NO. A simple model for cluster efficiency. In:
[22] Park J, Law KH. A data-driven, cooperative wind farm control to maximize the Proceedings of the European Wind Energy Association Conference and Exhibition;
total power production. Appl Energy 2016;165(Mar. 1):151–65. 1986. p. 407–10.
[23] Bastankhah M, Porté-agel F. A wind-tunnel investigation of wind-turbine wakes in [47] Voutsinas S, Rados K, Zervos A. On the analysis of wake effects in wind parks. Wind
different yawed and loading conditions. EGU General Assembly 2015. Engng 1990:204–219.
[24] Bastankhah M, Porté-agel F. Wind tunnel study of the wind turbine interaction [48] Hou Peng, Enevoldsen Peter, Weihao Hu, Chen Cong, Chen Zhe. Offshore wind
with a boundary-layer flow: Upwind region, turbine performance, and wake farm repowering optimization. Appl Energy, 2017;208(dec.15):834–44.
region. Phys Fluids 2017;29(6). [49] Segalini A, Jan-ke D. Blockage effects in wind farms. Wind Energy 2020;23.
[25] Bastankhah M, Porté-Agel F. Wind farm power optimization via yaw angle control: [50] Reddy SR. An efficient method for modeling terrain and complex terrain
A wind tunnel study. J Renewable Sustainable Energy 2019;11(2). boundaries in constrained wind farm layout optimization. Renewable Energy 2021;
[26] Gebraad PMO, Teeuwisse FW, Van Wingerden JW, Fleming PA, Ruben SD, 165:162–73.
Marden JR, et al. Wind plant power optimization through yaw control using a [51] Peña A, Réthoré P-E, Van Der Laan MP. On the application of the Jensen wake
parametric model for wake effects-a CFD simulation study. Wind Energy 2016;19 model using a turbulence-dependent wake decay coefficient: the Sexbierum case.
(1):95–114. Wind Energy 2015;19(4):763–76.
[27] Archer CL, Vasel-Be-Hagh A. Wake steering via yaw control in multi-turbine wind [52] Kragh KA, Hansen MH. Load alleviation of wind turbines by yaw misalignment.
farms: Recommendations based on large-eddy simulation. Sustainable Energy Wind Energy 2014;17(7):971–82.
Technol Assess 2019;33(JUN.):34–43. [53] Ali QS, Kim MH. Unsteady aerodynamic performance analysis of an airborne wind
[28] Howland MF, Lele SK, Dabiri JO. Wind farm power optimization through wake turbine under load varying conditions at high altitude. Energy Convers Manage
steering. PNAS 2019;116(29):14495–500. 2020;210:112696.
[29] Shapiro CR, Gayme DF, Meneveau C. Modelling yawed wind turbine wakes: a [54] Kanev SK, Savenije FJ. Active wake control: loads trends. Petten: ECN 2015.
lifting line approach. J Fluid Mech 2018;841:R1. [55] Zhang H, Ge M, Liu Y, Yang XI. A new coupled model for the equivalent roughness
[30] Coton FN, Wang T. The prediction of horizontal axis wind turbine performance in heights of wind farms. Renewable Energy 2021;171:34–46.
yawed flow using an unsteady prescribed wake model. Proc Inst Mech Eng Part A J [56] Ge M, Zhang S, Meng H, Ma H. Study on interaction between the wind-turbine
Power Energy 1999;213(1):33–43. wake and the urban district model by large eddy simulation. Renewable Energy
[31] Dou B, Guala M, Lei L, Zeng P. Wake model for horizontal-axis wind and 2020;157:941–50.
hydrokinetic turbines in yawed conditions. Appl Energy 2019;242(PT.1285-176): [57] Du B, Ge M, Zeng C, Cui G, Liu Y. Influence of atmospheric stability on wind-
1383–95. turbine wakes with a certain hub-height turbulence intensity. Phys Fluids 2021;33
[32] Lopez D, Kuo J, Li N. A novel wake model for yawed wind turbines. Energy 2019; (5):055111.
178(JUL.1):158–67. [58] Ge M, Gayme DF, Meneveau C. Large-eddy simulation of wind turbines immersed
in the wake of a cube-shaped building. Renewable Energy 2021;163:1063–77.

13

You might also like