Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Renewable Energy 197 (2022) 50–58

Contents lists available at ScienceDirect

Renewable Energy
journal homepage: www.elsevier.com/locate/renene

Wind farm blockage in a stable atmospheric boundary layer


Jessica M.I. Strickland, Srinidhi N. Gadde, Richard J.A.M. Stevens ∗
Physics of Fluids Group, Max Planck Center Twente for Complex Fluid Dynamics, J. M. Burgers Center for Fluid Dynamics, and MESA Research Institute,
University of Twente, P.O. Box 217, 7500 AE Enschede, The Netherlands

ARTICLE INFO ABSTRACT

Keywords: Wind farm blockage is the flow deceleration in front of a large turbine array, due to which the first wind farm
Wind energy row produces less power than a corresponding solitary row. Understanding wind farm blockage is crucial as
Wind farms a failure to accurately model the effect leads to a systematic difference between the modeled and actual wind
Wind farm blockage
farm performance. While stable atmospheric conditions are known to impact wind farm performance, the effect
Large-eddy simulation
on wind farm blockage is not well understood. We use large eddy simulations to investigate blockage in a stably
Stability
stratified atmospheric boundary layer without capping inversion by systematically varying the streamwise
turbine spacing and the surface cooling rate. We consider stable boundary layers without capping inversion as it
allows us to isolate the effect of surface stability. Primarily, we demonstrate that wind farm blockage increases
with atmospheric stability due to the deflection of relatively cold flow over the wind farm. Displacement of
this high-density colder air creates a high-pressure region at the wind farm entrance. The formation of the
high-pressure region enhances the adverse pressure gradient and increases the flow deceleration in front of
the wind farm under stable atmospheric conditions.

1. Introduction While the flow deceleration in front of a single turbine is well


understood, the influence of wind farm blockage, which is the flow
As the wind energy sector continues to grow, larger wind farms deceleration in front of a large turbine array, requires further study.
are designed and constructed while trying to minimize the turbine Wind farm blockage reduces the incoming kinetic energy available for
interactions which can negatively impact performance. Therefore, it is extraction by the first turbine row compared to the isolated turbine
essential to better understand the influence that closely spaced turbines case [14]. As a result, turbines in the first wind farm row produce less
have on each other. In general, to assess the effect of wind turbine power than they would in isolation, which has been confirmed in re-
wakes on the performance of downstream turbines, it is common to
cent field measurement campaigns [15–17]. Understanding wind farm
normalize the performance of turbines in a wind farm with the perfor-
blockage is particularly important as it leads to systematic changes in
mance of turbines in the first row [1,2]. This normalization implicitly
the estimated power production of a wind farm. In order to eventually
assumes that turbines in the first row operate in undisturbed inflow
include these effects in wind farm design models, it is first crucial to
conditions. However, wind turbine operation creates an induction re-
gion in front of the turbine, due to which turbines do not necessarily develop more fundamental understanding of the physics that governs
operate in undisturbed inflow conditions. For a single turbine, the wind farm blockage in stable atmospheric boundary layers (ABLs).
velocity deficit 𝑈 at hub-height and along the symmetry axis is well The ABL is the atmospheric layer directly in contact with Earth’s
described by the following analytical expression, which is obtained surface, which experiences a pronounced daily cycle [18]. The flow in
from actuator disk theory [3,4] or symmetric vortex theory [5], the ABL is turbulent, which promotes the vertical transport of momen-
[ ] tum, heat, and moisture. The turbulence develops because of vertical
𝑈 𝑥
=1−𝑎 1+ (1) wind shear (wind speed is zero at the ground) and buoyancy forces.
𝑈∞ (𝑥2 + 𝑅2 )1∕2
These processes are known as shear production and buoyancy produc-
where 𝑈∞ is the undisturbed upstream velocity, 𝑥 is the coordinate
tion/destruction of turbulence, respectively. The ABL characteristics
along the symmetry axis, 𝑅 is the rotor radius, and 𝑎 is the axial
change throughout the day. For example, during the daytime when the
induction factor. This expression has been validated using numerical
sun heats the surface, a convective (unstable) boundary layer forms,
models [6–9], and confirmed in wind tunnel experiments [10] and field
where buoyancy forces increase mixing and turbulence in the ABL.
measurements [11–13].

∗ Corresponding author.
E-mail address: r.j.a.m.stevens@utwente.nl (R.J.A.M. Stevens).

https://doi.org/10.1016/j.renene.2022.07.108
Received 2 February 2022; Received in revised form 3 July 2022; Accepted 20 July 2022
Available online 25 July 2022
0960-1481/© 2022 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
J.M.I. Strickland et al. Renewable Energy 197 (2022) 50–58

Fig. 1. Schematic depicting the (a) solitary infinite row and the (b) corresponding wind farm with eight rows. The inflow direction is indicated by the arrows and the dimensionless
spanwise and streamwise turbine spacing are 𝑠𝑦 = 𝑆𝑦 ∕𝐷 and 𝑠𝑥 = 𝑆𝑥 ∕𝐷, respectively.

When buoyancy forces do not directly affect the flow in the ABL, for Sebastiani et al. [15] used field measurements to show that strongly
instance around sunrise and sunset, in cloudy conditions, or offshore, stratified boundary layers increase wind farm blockage and highlight
the boundary layer is classified as neutral. At night, the surface cools, the challenges of assessing blockage with current industrial settings.
forming a stable boundary layer in which buoyancy forces destroy part Although the above studies agree that stability amplifies blockage,
of the shear-driven turbulence. Here, we study wind farm blockage in there is a discrepancy in the observed magnitude of the effect. And,
stable conditions because the increased mixing in a convective ABL is most importantly, the physical reason why atmospheric stability affects
expected to reduce blockage and most blockage studies to date have wind farm performance is not fully understood [14]. Here, we use
been performed in neutral conditions. As indicated below, wind farm LES to elucidate the effect of stratification on wind farm blockage.
blockage in stable atmospheric conditions is not fully understood yet. Specifically, we compare the performance of the first row of a wind
The potential impact of blockage on wind farm performance has farm to that of a solitary row (refer to Fig. 1) to isolate blockage effects
gained significant attention in recent years. Field measurements by caused by downstream turbines. We perform simulations of stable ABLs
Sebastiani et al. [15] found that wind farm blockage reduces the power without capping inversion as this eliminates the formation of gravity
production of the Lillgrund wind farm by about 2%. Nygaard et al. [17] waves [24–27] and allows us to isolate the effect of surface stability it-
introduced a wind farm blockage model based on the accumulation self. We will demonstrate that thermal stability exacerbates the adverse
of single-turbine induction effects. They find that their model cap- pressure gradient at the wind farm entrance, which increases wind farm
tures the blockage trends observed in field measurements performed blockage in stable conditions.
in the offshore wind farm Westernmost Rough, but underestimates the The remainder of this work is organized as follows. We first intro-
magnitude of the effect. Also, Branlard and Meyer Forsting [9] and duce the simulation framework and considered cases in Sections 2 and
Branlard et al. [19] used engineering-type models to account for wind 3. In Sections 4 and 5 we discuss the flow dynamics in ABLs and how
farm blockage. They showed that the flow velocity 2.5𝐷 upstream of a it affects wind farm performance, respectively. Subsequently, we will
5 × 5 wind farm with 5D inter turbine spacing could be reduced by 2%. discuss wind farm blockage in Section 6 and the physical mechanisms
Bleeg et al. [16] performed Reynolds Averaged Navier Stokes (RANS) behind it in Section 7. The conclusions are provided in Section 8.
simulations and observed, in reasonable agreement with their field
observations, that wind farm blockage causes a wind-speed reduction of 2. Large eddy simulations
1.9% at 7−10𝐷 upstream. Later, Bleeg [20] suggested employing RANS
models in combination with a graph neural network approach to model We use LES to investigate wind farm blockage in a stable ABL
turbine interaction losses and blockage effects because traditional wake using a code that originates from the work by Albertson and Par-
models tend to ignore blockage effects. lange [28,29]. The code has been successfully used to simulate wind
Segalini et al. [21,22] used wind tunnel experiments with uniform farms in stably stratified ABL [30,31] and is described in detail below.
inflow to study wind farm blockage. They established an empirical The governing equations are the filtered Navier–Stokes equations with
relationship that captures the effect of wind farm blockage on the buoyancy modeled using the Boussinesq approximation
performance of the first wind farm row. Segalini [14] introduced an ( )
analytical model based on the linearized RANS equations, which was 𝑢 𝑖 + 𝜕𝑗 ̃
𝜕𝑡 ̃ 𝑢𝑗 = − 𝜕𝑖 𝑝̃ − 𝜕𝑗 𝜏𝑖𝑗 + 𝑔𝛽(𝜃̃ − 𝜃̃0 )𝛿𝑖3
𝑢𝑖 ̃
(2)
validated against their wind tunnel measurement data. They also find + 𝑓𝑐 (𝑈𝑔 − ̃ 𝑢)𝛿𝑖2 − 𝑓𝑐 (𝑉𝑔 − 𝑣̃)𝛿𝑖1 + 𝑓̃𝑥 𝛿𝑖1 + 𝑓̃𝑦 𝛿𝑖2 ,
good agreement between model and measurement data, although they
point out that the model has some deficiencies. For example, the 𝑢𝑖 = 0 ,
𝜕𝑖 ̃ (3)
model does not predict that a single row farm experiences blockage.
Large eddy simulations (LES) [23] revealed that trends observed by 𝜕𝑡 𝜃̃ + ̃
𝑢𝑗 𝜕𝑗 𝜃̃ = − 𝜕𝑗 𝑞𝑗 , (4)
the empirical relation by Segalini et al. [21,22] are representative of
where the tilde ( ̃ ) represents spatial filtering at the resolved scale
the trends observed in neutral ABLs. ( )
̃ and ̃
𝛥, 𝑢𝑖 = ̃ ̃ is the velocity, 𝑝̃ = 𝑝̃∗ ∕𝜌 + 𝜎𝑘𝑘 ∕3 represents
𝑢, 𝑣̃, 𝑤
Although it is well known that wind farm performance is affected
the modified pressure by adding the trace of sub-grid stress 𝜎𝑘𝑘 ∕3 to
by atmospheric stability [24–27], the effect of stable atmospheric strat-
the kinematic pressure 𝑝̃∗ ∕𝜌, and 𝜌 is the density. The buoyancy term
ification on wind farm blockage is relatively unexplored [14]. Some ̃ the acceleration due to gravity 𝑔,
contains the potential temperature 𝜃,
⟨ ⟩
studies in conventionally neutral ABLs attribute wind farm blockage
and the buoyancy parameter 𝛽 = 1∕ 𝜃̃ , where the planar averaged po-
to the pressure gradient induced by gravity waves [24–27]. However, ⟨ ⟩
gravity waves make it complicated to isolate the effect of surface tential temperature 𝜃̃ is used as the reference temperature, and 𝛿𝑖𝑗 is
stability on wind farm blockage. Simley et al. [13] showed that sta- the Kronecker delta function. The Coriolis forcing is represented in two
bility may exaggerate the induction preceding a single turbine, and components accompanied by the Coriolis parameter 𝑓𝑐 = 2𝛺 sin 𝜙 =

51
J.M.I. Strickland et al. Renewable Energy 197 (2022) 50–58

1.159 × 10−4 rad∕s where 𝛺 is the rotation of the Earth and 𝜙 = 52.8◦ , placed at 𝑥 = 6 km, which is more than sufficient to properly capture
corresponding to the Dutch North Sea. The flow is driven by a mean the induction region [48].
pressure gradient 𝑝∞ that is represented by the geostrophic wind in The simulations are carried out in two stages. In the initial stage,
𝜕𝑝 𝜕𝑝
which 𝑈𝑔 = − 𝜌𝑓1 𝜕𝑦∞ and 𝑉𝑔 = 𝜌𝑓1 𝜕𝑥∞ are the streamwise and spanwise a stable ABL simulation is performed for approximately 7 h until a
𝑐 𝑐
components, respectively. The geostrophic wind speed magnitude 𝐺 =
√ quasi-steady state is achieved. In the second stage, the concurrent
𝑈𝑔2 + 𝑉𝑔2 = 12 m∕s. precursor method is applied [49], during which data from the stable
ABL simulation is sampled and used as an inflow condition for a
The turbine forces 𝑓̃𝑖 = (𝑓̃𝑥 , 𝑓̃𝑦 , 0) are modeled with the filtered
second domain in which the turbines are placed. The simulations are
actuator disk approach [32–35]. The sub-grid scale terms 𝜏𝑖𝑗 and 𝑞𝑗 ,
performed long enough to ensure that the power production of the wind
which represent the traceless part of the sub-grid scale stress tensor
farm has reached its quasi-stationary state. We collect statistics between
and the sub-grid scale heat flux tensor, respectively, are modeled using
hour 8 and 9 of the simulation.
a tuning-free, scale-dependent, dynamic anisotropic minimum dissipa-
tion model [31,36–38]. We perform wall-modeled LES, and therefore
4. Boundary layer characteristics
the wall-stress and fluxes are modeled using the Monin–Obukhov sim-
ilarity theory [39]. The governing equations are discretized using a
In Fig. 2, we first investigate the characteristics of the various
pseudo-spectral method in the horizontal directions and central differ-
ABL cases considered here. Here, and in the remainder of this work,
ences in the vertical direction. More information about the numerical
horizontal-averaging is denoted by angular brackets (⟨ ⟩) and time-
method used in the code can be found in Albertson [28] and Gadde and
averaging is denoted by a bar ( ̄ ). Fig. 2(a) reveals that the potential
Stevens [27]. The code has been validated for stable boundary layer
temperature is impacted by the surface cooling rate and increases
simulations in Gadde et al. [31]. Furthermore, the code has been val-
with height for the stable cases, while it remains constant for the
idated against wind tunnel experiments, which characterized the flow neutral case. Fig. 2(b)
in a model wind farm in detail, by Chamorro and Porté-Agel [40,41] √ shows the time and horizontally-averaged wind
magnitude 𝑣mag = 𝑢̄ 2 + 𝑣̄ 2 as a function of height. There is a clear
and Stevens et al. [42].
distinction between the neutral and stable cases. In particular, for the
stable cases, a low-level jet is formed [50]. We note that the low-
3. Simulation setup level jet becomes more intense and closer to the ground when the
stability increases. For instance, when 𝐶𝑟 = 1.0 K∕hr, the low-level jet
We perform LES of stable ABLs by systematically reducing the height is within the rotor-swept area of the turbines. Due to its high
cooling rate 𝐶𝑟 at the surface, so that the air is relatively cooler than kinetic energy and strong shear, the low-level jet plays an essential
the air above it, which inhibits air movement. Typically, LES studies of role in sustaining turbulence within the ABL, and therefore these jets
wind farm performance in stable boundary layers focus on cases with have a pronounced influence on wind farm performance [27]. Fig. 2(c)
a strong capping inversion representing a stable stratification above presents the turbulence intensity profiles, which are defined as 𝐼 =
the boundary layer. In such cases, it is difficult to isolate the impact 𝜎∕𝑣mag with 𝜎 = 𝑢′2 + 𝑣′2 divided by the local wind velocity magnitude.
of stable stratification itself on wind farm blockage. Therefore, we The figure reveals that as the surface cooling rate increases, the ABL
do not impose a capping inversion because this is known to trigger becomes more stably stratified and there is less mixing. As the mixing
gravity waves, which would ultimately affect the flow within the 2 2
changes, so does the vertical momentum flux 𝜏 = (𝑢′ 𝑤′ + 𝑣′ 𝑤′ )1∕2 as
domain [24,25,43–46]. This approach allows us to isolate the effect of shown in Fig. 2(d). These boundary layer characteristics influence the
surface cooling on flow blockage. momentum exchange in the boundary layer and the interaction with
We perform 20 LES to investigate the impact of the streamwise the wind farm. Fig. 2(e) shows the wind direction (𝛼 = tan−1 (𝑣∕𝑢)) as
turbine spacing 𝑆𝑥 and the surface cooling rate 𝐶𝑟 on wind farm block- a function of height. This figure reveals that the wind veer is strongly
age. We consider four different surface cooling rates (𝐶𝑟 = 0, 0.25, 0.5, affected by the atmospheric stability and is strongest for the most stable
1.0 K∕hr). When the cooling rate at the ground is set to 𝐶𝑟 = 0 K∕hr, case with a cooling rate of 𝐶𝑟 = 1.0 K∕hr.
a neutral ABL is generated. Gadde et al. [27] confirmed that this
method results in negligible surface heat flux 𝑞∗ . Furthermore, we select 5. Wind farm performance
three typical cooling rates at the surface to adjust the level of thermal
stratification in the surface layer of the ABL. The atmospheric strati- Fig. 3(a,b) display the instantaneous wind velocity magnitude at
fication increases when the surface cooling rate is increased, which is hub-height for a wind farm with 𝑆𝑥 = 7𝐷 in neutral and stable ABL
the nomenclature we will use going forward. conditions. In both cases pronounced velocity deficits are visible behind
For each cooling rate, we consider a reference solitary row case each turbine as momentum is extracted from the wind. However, the
and four different wind farm layouts as shown in Fig. 1. Each wind shape and magnitude of the wakes depend strongly on atmospheric sta-
farm layout has 8 rows in downstream direction and the streamwise bility. The higher turbulence intensity for the neutral case is reflected
turbine spacing is varied (𝑆𝑥 ∕𝐷 = 7, 5, 3.5, 1.75). Each row has eight in the larger and more prominent turbulent structures. Furthermore,
turbines and the dimensionless spanwise spacing (𝑆𝑦 ∕𝐷 = 4.608) re- the wake recovery is slower in stable conditions compared to neutral
mains constant. The domain is spanwise periodic; hence, we effectively conditions because the reduced turbulent mixing limits the entrainment
consider infinitely long rows of turbines. We have previously shown of high energetic flow from above the wind farm. The corresponding
that preventing air from flowing around the wind farm increases the time-averaged fields in Fig. 3(c,d) more clearly reveal the clockwise
magnitude of the observed wind farm blockage (see also the work deflection of the wind turbine wakes that are observed towards the end
by Bleeg and Montavon [47]), but does not fundamentally alter the of the wind farm in stable conditions [30,51]. Furthermore, comparing
blockage effect induced by downstream turbines [23]. Figs. 3(e) and (f) reveal that the strong wind veer in stable conditions
The following parameters of the wind farm setup remain constant deflects the wakes at altitudes close to the turbine top, due to which
throughout this study: the turbine diameter is 𝐷 = 125 m, the hub- turbines in the second and third-row are experiencing less wake effects.
height is 𝐻𝑇 = 100 m, and the thrust coefficient is 𝐶𝑇 = 0.85. The This effect is not observed in neutral conditions where wind veer is
turbine-arrays are placed in a computational domain of 𝐿𝑥 × 𝐿𝑦 × 𝐿𝑧 = limited.
17.28 km × 4.608 km × 8.6 km, discretized by 𝑁𝑥 × 𝑁𝑦 × 𝑁𝑧 = 1440 × Both thermal stratification and the wind farm layout affect wind
384×384 grid points. Therefore, the corresponding horizontal resolution farm power production. Fig. 4 presents the average power production
is 𝛥𝑥 = 𝛥𝑦 = 12 m. The vertical resolution up to 𝑧 = 1000 m is of each turbine row within a wind farm normalized by the production
𝛥𝑧 = 5 m, above which the grid is stretched. The first turbine row is of the first row. The figure confirms that downstream rows produce

52
J.M.I. Strickland et al. Renewable Energy 197 (2022) 50–58

Fig. 2. Time and horizontally-averaged (a) potential temperature, (b) velocity magnitude, (c) turbulence intensity, (d) vertical momentum flux, and (e) wind direction as function
of height for the different stability cases, indicated by the cooling rates 𝐶𝑟 .

Fig. 3. Instantaneous velocity magnitude at hub-height normalized with the incoming velocity at hub-height in a wind farm with 𝑆𝑥 = 7𝐷 in a (a) neutral (𝐶𝑟 = 0.0 K∕hr) and (b)
stable (𝐶𝑟 = 1.0 K∕hr) ABL. (c,d) Corresponding time-averaged velocities at hub-height, and (e,f) close to the top of the turbine.

53
J.M.I. Strickland et al. Renewable Energy 197 (2022) 50–58

Fig. 4. The power production of each row in the wind farm (Pf ) normalized by the first row production of each wind farm (Pf (𝑟𝑜𝑤=1) ) for wind farms with different streamwise
turbine spacing.

Fig. 5. (a) The normalized power production of the first turbine row as a function of time for wind farms with 𝑆𝑥 = 5𝐷 and different thermal stabilities. (b) Standard deviation
of the normalized first row power production.

much less than the first row due to wakes and consequently, the depend on the atmospheric conditions; however, there is no significant
relative power production of downstream rows strongly depends on dependence on the streamwise spacing of the wind farm.
the streamwise turbine spacing. For most cases, we observe that the We emphasize that the results in Fig. 4 only provide a measure of
relative performance of each downstream row reaches a constant level the balance between wake losses and flow recovery inside the wind
as the kinetic energy extracted by upstream turbines is replenished farm. However, this classical representation in which the results are
through vertical kinetic energy entrainment as the flow progresses normalized with the first-row performance does not allow one to study
downstream [1,2,33,52]. Fig. 4(a) shows that for the most stable case wind farm blockage. Therefore, in the next section, we compare the
(𝐶𝑟 = 1.0 K∕hr) and 𝑆𝑥 = 7𝐷 the observed trend is different. This performance of the first wind farm row with the performance of a
behavior is due to the strong wind veer in stable conditions, which solitary row.
can lead to significant wake deflection at certain heights. As shown in
Fig. 3(f), the deflection close to the top of the turbine causes turbines in
6. Wind farm blockage
the second and third-row to produce more. However, this reduces the
incoming flow energy available for turbines further downstream. At the
same time, the vertical kinetic energy entrainment is more limited for So far, by normalizing the result with the first wind farm row, we
cases with higher thermal stability (see the energy budget analysis in assumed that the first row operates in undisturbed inflow conditions.
Section 7), which limits the power production of turbines in the back However, this assumption is not generally valid. Therefore we compare
of the wind farm. the production of the first wind farm row to the production of a solitary
In Fig. 5, we show the time-dependent power production of the first row to study wind farm blockage [14,16,22,23]. Fig. 6 shows the time-
row for the 𝑆𝑥 = 5𝐷 cases for the last hour of the simulation. The figure averaged power production of first wind farm row normalized by the
reveals that the first-row power fluctuations decrease with atmospheric corresponding solitary row for all cases. Primarily, we observe that
stability due to the reduced turbulence. The gradual increase in the the power production of the first wind farm row is lower than that of
power production over time originates from the inertial oscillation in the solitary row for all cases. Fig. 6(a) shows that wind farm blockage
the velocity profile in the boundary layer due to the Coriolis forces. increases with atmospheric stability. For the neutral case the observed
This makes the mean velocity profile to continuously change over time blockage increases from ∼ 1% for 𝑆𝑥 = 7𝐷 to ∼ 7% for 𝑆𝑥 = 1.75𝐷,
with a time period of 15 h. The standard deviation of the normalized while for the most stable case it increases from ∼ 4% for 𝑆𝑥 = 7𝐷 to
time-dependent power production of the first row for all cases is pre- ∼ 15% for 𝑆𝑥 = 1.75𝐷. Fig. 6(b) emphasizes that the blockage indeed
sented in Fig. 5(b). This figure confirms that these fluctuations strongly strongly depends on the streamwise turbine spacing and that for a given

54
J.M.I. Strickland et al. Renewable Energy 197 (2022) 50–58

Fig. 6. The average power production of a turbine in the first wind farm row (Pf (𝑟𝑜𝑤=1) ) compared to the power production of a corresponding solitary row Pr as a function of (a)
the cooling rate and (b) dimensionless streamwise turbine spacing.

power production in different parts of the wind farm [27,44], especially


in the entrance region of the wind farm where the blockage is observed.
The budget equation is obtained by multiplying the momentum Eq. (2)
by 𝑢𝑖 . The different energy terms are then integrated over the control
𝐴
volume surrounding each turbine row. We select control volumes
of size 𝑆𝑥 × 𝐿𝑦 × 𝐷 surrounding the center of each turbine row. The
corresponding energy equation is given by:
( ) ( )
1 1 1 ′ ′ ′ 1 ′ ′
𝑢𝑖 𝑢𝑗 + 𝑢′𝑖 𝑢′𝑗 𝑑𝑆𝑖 +
𝐴
𝑓𝑖 𝑢𝑖 𝑑 = 𝑢𝑗 𝑢𝑗 𝑢𝑖 𝑢𝑖 + 𝑢𝑖 𝑢𝑖 𝑢𝑗 𝑑𝑆𝑖
∫ 𝐴 ∫𝑆 2 2 ∫𝑆 2 2
⏟⏞⏞⏞⏟⏞⏞⏞⏟ ⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟ ⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
P, Turbine power Ek , Kinetic energy flux Tt , Turbulent transport
( ) ( )
+ 𝑢𝑖 𝜏𝑖𝑗 𝑑𝑆𝑖 + 𝑝𝑢𝑖 𝑑𝑆𝑖
∫𝑆 ∫𝑆
⏟⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏟ ⏟⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏟
Tsgs , SGS transport F, Flow work
( ) ( )
𝑔𝛽𝛿𝑖3 𝑢𝑖 𝜃 − 𝑢𝑖 ⟨𝜃⟩ 𝑑 −
𝐴 𝐴
− 𝜏𝑖𝑗 𝑆𝑖𝑗 𝑑
∫ 𝐴 ∫ 𝐴
Fig. 7. The time and spanwise-averaged vertical velocity as a function of the upstream ⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟ ⏟⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏟
position at hub-height for three different turbine arrays and neutral (𝐶𝑟 = 0.0 K∕hr) B, Buoyancy D, Dissipation
and stable (𝐶𝑟 = 1.0 K∕hr) conditions as indicated in the legend. ( ) 𝐴
− 𝑓𝑐 𝑢𝑖 𝑈𝑔 𝛿𝑖2 − 𝑓𝑐 𝑢𝑖 𝑉𝑔 𝛿𝑖1 𝑑 . (5)
∫ 𝐴
⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏞⏟
G, Geostrophic forcing
streamwise turbine spacing, the observed wind farm blockage increases
with atmospheric stability. In the above equation P is the time-averaged power per turbine row
To better understand wind farm blockage, we analyze how the flow and Ek is the mean-flow transport of kinetic energy. Tt and Tsgs are the
in front of the wind farm is affected. Fig. 7 shows the normalized time transport of momentum by resolved and SGS turbulent fluxes, respec-
and spanwise-averaged vertical velocity at hub-height for the solitary tively. The transfer of energy due to pressure changes is represented by
row (𝑆𝑥 = 0) and two wind farm layouts (𝑆𝑥 = 5𝐷, 1.75𝐷) for neutral F, while B is the turbulence destruction or production due to buoyancy,
and stable ABL conditions. We find that the positive vertical velocity in and G is the mean geostrophic forcing.
the induction region preceding the turbine array increases as the flow Fig. 9 shows the energy budget for the wind farms with 𝑆𝑥 = 7𝐷
approaches the wind farm. This indicates that the flow deflection over and different thermal stratification. The figure reveals that the power
the wind farm increases with increasing atmospheric stratification and production of the first row is dominated entirely by the mean kinetic
when the streamwise turbine spacing is smaller. energy Ek and flow work F. In contrast, the turbulent transport T𝑡 ,
Fig. 8 compares the normalized streamwise velocity at hub-height dissipation D, and buoyancy B are negligible. Therefore, blockage is
along the centerline preceding each turbine in the induction region for a non-turbulent process that depends solely on the mechanical energy
the same wind farm layouts. Specifically, the deceleration in front of in the flow. This observation coincides with the observations based on
the solitary row (𝑆𝑥 = 0) and two wind farm layouts with different RANS studies elsewhere [14,22,47].
streamwise spacing (𝑆𝑥 = 1.75𝐷, 5𝐷) is compared. The induction in Further downstream in the wind farm, the turbulent entrainment
front of an isolated turbine is provided as a reference; see Eq. (1). In T𝑡 increases due to the wind turbine wakes, while the mean kinetic
agreement with the flow deflection over the wind farm, Fig. 8 reveals energy Ek gradually decreases. The other terms are significantly smaller
that flow deceleration in front of the turbine array increases with throughout the entire wind farm, although D slightly increases inside
atmospheric stability and reduced streamwise turbine spacing. the wind farm due to the wind turbine wakes. The figure shows that at
the back of the wind farm the power production is essentially balanced
7. Flow analysis of wind farm blockage by the turbulent entrainment T𝑡 [1,2,33,52,53]. Furthermore, Fig. 9
shows that the turbulent entrainment reduces when the atmospheric
It is essential to understand how the presence of stable stratification stability is increased, which explains the lower power production at the
increases the upstream flow deceleration. Therefore, we perform an en- end of the wind farm for the most stable case (𝐶𝑟 = 1.0 K∕hr). For this
ergy budget analysis to get insight into the processes that dominate the case, the mean kinetic energy Ek is significantly higher in the entrance

55
J.M.I. Strickland et al. Renewable Energy 197 (2022) 50–58

Fig. 8. Time-averaged streamwise velocity at hub-height along the centerline preceding each turbine normalized by the incoming wind speed at hub-height. The line types and
colors indicates the result for different streamwise turbine spacings and thermal stabilities as indicated in the legend. The purple line is the reference for a stand-alone turbine,
see Eq. (1).

Fig. 9. Energy budget analysis for wind farms with 𝑆𝑥 = 7𝐷 in different stability conditions.

region of the farm than for the other cases. The reason for this behavior Fig. 11 shows the time and spanwise averaged buoyancy forces cal-
is the wake deflection, due to which turbines on the second and third culated based on the Boussinesq approximation under stable conditions
row can produce more, as also shown in Fig. 3.
for the wind farm case with 𝑠𝑥 = 1.75𝐷. Negative values indicate that
A crucial finding from the energy budget analysis is that the pres-
the air temperature is lower than the average value at that height. The
sure drop and mean kinetic energy Ek are the dominant processes that
determine the power production of turbines on the first row. Therefore, figure shows that cold air is deflected over the wind farm. According
we analyze the pressure distribution around the wind farm in more to the hydrostatic law [46], the pressure decreases with height and is
detail. Fig. 10(a) shows the pressure distribution in and around the equal to the weight of the fluid above, i.e. 𝑑𝑃 ∕𝑑𝑧 = −𝜌𝑔. Essentially,
wind farm under neutral and stable conditions. This figure shows the
the pressure at a particular height depends on the density of the fluid
adverse pressure gradient (positive 𝑑𝑃 ∕𝑑𝑥) in front of the wind farm,
at that height, and therefore, the pressure is highest at the ground
which leads to the above-observed flow deceleration in the induction
region (Fig. 8). In the Navier–Stokes equation, the acceleration terms and decreases with height. The relatively cold air deflected over the
can be equated to the forces acting on the fluid domain, and accordingly wind farm increases the pressure at the wind farm entrance, which
𝑑𝑢∕𝑑𝑡 ≈ −𝐷𝑃 ∕𝑑𝑥, which means that the adverse pressure gradient increases the streamwise adverse pressure gradient and consequently,
leads to a flow deceleration in front of the wind farm. Fig. 10(b) shows
flow deceleration preceding the wind farm. Therefore, the enhanced
that the adverse pressure gradient at hub-height is stronger under sta-
adverse pressure gradient caused by the deflection of cold flow over
ble stratification than under neutral stratification. To understand why
stable stratification increases the magnitude of the adverse pressure the wind farm is responsible for increased wind farm blockage with
gradient, we look at the relative thermal stability inside the wind farm. stronger thermal stratification.

56
J.M.I. Strickland et al. Renewable Energy 197 (2022) 50–58

Fig. 10. (a) Spanwise-averaged pressure (𝑃 ∕𝜌 [m∕s2 ]) under neutral (𝐶𝑟 = 0.0 K∕hr) and stable (𝐶𝑟 = 1.0 K∕hr) stratification for the 𝑆𝑥 = 1.75𝐷 wind farm layout, and (b)
streamwise pressure, normalized with the inlet pressure, at hub-height.

Fig. 11. Spanwise-averaged buoyancy force under stable (𝐶𝑟 = 1.0 K∕hr) stratification where 𝑠𝑥 = 1.75𝐷.

8. Conclusions Data availability statement

We use LES to show that wind farm blockage increases with ther- The data that support the findings of this study are available from
mal stability and reduced streamwise turbine spacing. In general, we the corresponding author upon reasonable request.
observed that wind farm productivity decreases with increased at-
mospheric stability and reduced streamwise turbine spacing due to Acknowledgments
decreased entrainment and increased turbine interactions. Wind farm
blockage is known to reduce the power production of the first wind This work is part of the Shell-NWO/FOM-initiative Computational
farm row when downstream turbines are placed close by. We showed sciences for energy research of Shell and Chemical Sciences, Earth and
that wind farm blockage increases with atmospheric stability. Further- Live Sciences, Physical Sciences, FOM, and STW. This work was partly
more, an energy budget analysis reveals that the performance of the carried out on the national e-infrastructure of SURFsara, a subsidiary
first wind farm row is dominated entirely by the mean kinetic energy of SURF corporation, the collaborative ICT organization for Dutch
and flow work, i.e. the transfer of energy due to pressure changes. Sub- education and research, and an STW VIDI grant (No. 14868). We
sequently, we demonstrate that higher-density, cold air moves up due acknowledge PRACE for awarding us access to MareNostrum 4 based in
to the wind farm blockage in a stable boundary layer. This deflection Spain at the Barcelona Computing Center (BSC) under PRACE project
2020225335 and 2020235589.
of cold air over the wind farm creates a cold anomaly above the wind
farm that increases pressure at the wind farm entrance according to the
hydrostatic law [46]. The increased pressure at the wind farm entrance References
amplifies the adverse pressure gradient which decelerates the flow in
[1] R.J.A.M. Stevens, C. Meneveau, Flow structure and turbulence in wind farms,
the region in front of the wind farm and explains the increased wind Annu. Rev. Fluid Mech. 49 (2017) 311–339, http://dx.doi.org/10.1146/annurev-
farm blockage with thermal stratification. fluid-010816-060206.
[2] F. Porté-Agel, M. Bastankhah, S. Shamsoddin, Wind-turbine and wind-farm flows:
A review, Boundary-Layer Meteorol. 74 (2020) 1–59, http://dx.doi.org/10.1007/
CRediT authorship contribution statement s10546-019-00473-0.
[3] C. Koning, Influence of the propeller on other parts of the airplane structure, in:
Aerodynamic Theory, Springer, 1935, pp. 361–430, http://dx.doi.org/10.1007/
Jessica M.I. Strickland: Conceptualization, Formal analysis, Com- 978-3-642-91487-4_4.
putation, Writing – original draft, Writing – review & editing. Srinidhi [4] H. Glauert, Airplane Propellers Aerodynamic Theory Volume IV Edited By
N. Gadde: Conceptualization, Formal analysis, Software, Methodol- William Frederick Durand, The Dover edition, 1963.
ogy, Validation, Writing – review & editing. Richard J.A.M. Stevens: [5] N. Joukowsky, Vortex theory of screw propeller, in: I Tr. Otd. Fiz. Nauk O-Va
Lubitelei Estestvozn, Vol. 16, 1912, pp. 1–31.
Conceptualization, Funding acquisition, Resources, Writing – review &
[6] A.R. Meyer Forsting, N. Troldborg, J.P.M. Leon, A. Sathe, N. Angelou, A.
editing, Supervision. Vignaroli, Validation of a CFD model with a synchronized triple-lidar system
in the wind turbine induction zone, Wind Energy 20 (2017) 1481–1498, http:
//dx.doi.org/10.1002/we.2103.
Declaration of competing interest [7] A.R. Meyer Forsting, M.P. van der Laan, N. Troldborg, The induction zone/factor
and sheared inflow: A linear connection? J Phys. Conf. Ser. 1037 (2018) 072031,
http://dx.doi.org/10.1088/1742-6596/1037/7/072031.
The authors declare that they have no known competing finan-
[8] N. Troldborg, A.R. Meyer Forsting, A simple model of the wind turbine
cial interests or personal relationships that could have appeared to induction zone derived from numerical simulations, Wind Energy 20 (12) (2017)
influence the work reported in this paper. 2011–2020, http://dx.doi.org/10.1002/we.2137.

57
J.M.I. Strickland et al. Renewable Energy 197 (2022) 50–58

[9] E. Branlard, A.R. Meyer Forsting, Assessing the blockage effect of wind turbines [33] M. Calaf, C. Meneveau, J. Meyers, Large eddy simulations of fully developed
and wind farms using an analytical vortex model, Wind Energy 23 (2020) wind-turbine array boundary layers, Phys. Fluids 22 (2010) 015110, http://dx.
2068–2086, http://dx.doi.org/10.1002/we.2546. doi.org/10.1063/1.3291077.
[10] D. Medici, S. Ivanell, J.-Å. Dahlberg, P.H. Alfredsson, The upstream flow of [34] M. Calaf, M.B. Parlange, C. Meneveau, Large eddy simulation study of scalar
a wind turbine: blockage effect, Wind Energy 14 (5) (2011) 691–697, http: transport in fully developed wind-turbine array boundary layers, Phys. Fluids 23
//dx.doi.org/10.1002/we.451. (2011) 126603, http://dx.doi.org/10.1063/1.3663376.
[11] E. Simley, L.Y. Pao, P. Gebraad, M. Churchfield, Investigation of the impact of [35] C.R. Shapiro, D.F. Gayme, C. Meneveau, Filtered actuator disks: Theory and
the upstream induction zone on LIDAR measurement accuracy for wind turbine application to wind turbine models in large eddy simulation, Wind Energy (2019)
control applications using large-eddy simulation, J. Phys. Conf. Ser. 524 (1) 1–7, http://dx.doi.org/10.1002/we.2376.
(2014) 012003, http://dx.doi.org/10.1088/1742-6596/524/1/012003. [36] R. Verstappen, When does eddy viscosity damp subfilter scales sufficiently? J.
[12] K.B. Howard, M. Guala, Upwind preview to a horizontal axis wind turbine: Sci. Comp. 49 (1) (2011) 94, http://dx.doi.org/10.1007/s10915-011-9504-4.
a wind tunnel and field-scale study, Wind Energy 19 (8) (2016) 1371–1389, [37] M. Abkar, A. Sharifi, F. Porté-Agel, Wake flow in a wind farm during a diurnal
http://dx.doi.org/10.1002/we.1901. cycle, J. Turb. 17 (4) (2016) 420–441, http://dx.doi.org/10.1080/14685248.
[13] E. Simley, N. Angelou, T. Mikkelsen, M. Sjöholm, J. Mann, L.Y. Pao, Charac- 2015.1127379.
terization of wind velocities in the upstream induction zone of a wind turbine [38] M. Abkar, P. Moin, Large eddy simulation of thermally stratified atmospheric
using scanning continuous-wave lidars, J. Renew. Sustain. Energy 8 (1) (2016) boundary layer flow using a minimum dissipation model, Boundary-Layer Me-
013301, http://dx.doi.org/10.1063/1.4940025. teorol. 165 (3) (2017) 405–419, http://dx.doi.org/10.1007/s10546-017-0288-
[14] A. Segalini, An analytical model of wind-farm blockage, J. Renew. Sustain. 4.
Energy 13 (2021) 033307, http://dx.doi.org/10.1063/5.0046680. [39] C.-H. Moeng, A large-eddy simulation model for the study of planetary boundary-
[15] A. Sebastiani, F. Castellani, G. Crasto, A. Segalini, Data analysis and simulation layer turbulence, J. Atmos. Sci. 41 (1984) 2052–2062, http://dx.doi.org/10.
of the Lillgrund wind farm, Wind Energy 24 (6) (2021) 634–648, http://dx.doi. 1175/1520-0469(1984)041%3C2052:ALESMF%3E2.0.CO;2.
org/10.1002/we.2594. [40] L.P. Chamorro, F. Porté-Agel, Effects of thermal stability and incoming
[16] J. Bleeg, M. Purcell, R. Ruisi, E. Traiger, Wind farm blockage and the conse- boundary-layer flow characteristics on wind-turbine wakes: A wind-tunnel
quences of neglecting its impact on energy production, Energies 11 (6) (2018) study, Boundary-Layer Meteorol. 136 (2010) 515–533, http://dx.doi.org/10.
1609, http://dx.doi.org/10.3390/en11061609. 1007/s10546-010-9512-1.
[17] N.G. Nygaard, S.T. Steen, L. Poulsen, J.G. Pedersen, Modeling cluster wakes and [41] L.P. Chamorro, F. Porté-Agel, Turbulent flow inside and above a wind farm:
wind farm blockage, J. Phys. Conf. Ser. 1618 (2020) 062072, http://dx.doi.org/ a wind-tunnel study, Energies 4 (2011) 1916–1936, http://dx.doi.org/10.3390/
10.1088/1742-6596/1618/6/062072. en4111916.
[18] R.B. Stull, An Introduction to Boundary Layer Meteorology, Kluwer Academic [42] R.J.A.M. Stevens, L.A. Martínez-Tossas, C. Meneveau, Comparison of wind farm
publishers, Boston, 1988, http://dx.doi.org/10.1007/978-94-009-3027-8. large eddy simulations using actuator disk and actuator line models with wind
[19] E. Branlard, E. Quon, A.R. Meyer Forsting, J. King, P. Moriarty, Wind farm tunnel experiments, Renew. Energy 116 (2018) 470–478, http://dx.doi.org/10.
blockage effects: comparison of different engineering models, J. Phys. Conf. Ser. 1016/j.renene.2017.08.072.
1618 (2020) 062054, http://dx.doi.org/10.1088/1742-6596/1618/6/062036. [43] D.G. Dommermuth, D.K.P. Yue, A high-order spectral method for the study of
[20] J. Bleeg, A graph neural network surrogate model for the prediction of turbine nonlinear gravity waves, J. Fluid Mech. 184 (1987) 267–288, http://dx.doi.org/
interaction loss, J. Phys. Conf. Ser. 1618 (2020) 062054, http://dx.doi.org/10. 10.1017/S002211208700288X.
1088/1742-6596/1618/6/062054. [44] D. Allaerts, J. Meyers, Boundary-layer development and gravity waves in
[21] A. Segalini, J.-Å. Dahlberg, Global blockage effects in wind farms, J. Phys. conventionally neutral wind farms, J. Fluid Mech. 814 (2017) 95–130, http:
Conf. Ser. 1256 (1) (2019) 012021, http://dx.doi.org/10.1088/1742-6596/1256/ //dx.doi.org/10.1017/jfm.2017.11.
1/012021. [45] D. Allaerts, J. Meyers, Gravity waves and wind-farm efficiency in neutral and
[22] A. Segalini, J.-Å. Dahlberg, Blockage effects in wind farms, Wind Energy 23 (2) stable conditions, Boundary-Layer Meteorol. 166 (2018) 269, http://dx.doi.org/
(2020) 120–128, http://dx.doi.org/10.1002/we.2413. 10.1007/s10546-017-0307-5.
[23] J.M.I. Strickl, R.J.A.M. Stevens, Investigating wind farm blockage in a neutral [46] R.B. Smith, Gravity wave effects on wind farm efficiency, Wind Energy 13 (5)
boundary layer using large-eddy simulations, Eur. J. Mech. B/Fluids 95 (2022) (2010) 449–458, http://dx.doi.org/10.1002/we.366.
303–314, http://dx.doi.org/10.1016/j.euromechflu.2022.05.004. [47] J. Bleeg, C. Montavon, Blockage effects in a single row of wind turbines,
[24] K.L. Wu, F. Porté-Agel, Flow adjustment inside and around large finite-size wind J. Phys. Conf. Ser. 2265 (2) (2022) 022001, http://dx.doi.org/10.1088/1742-
farms, Energies 10 (12) (2017) 2164, http://dx.doi.org/10.3390/en10122164. 6596/2265/2/022001.
[25] D. Allaerts, J. Meyers, Sensitivity and feedback of wind-farm-induced gravity [48] J.M.I. Strickl, R.J.A.M. Stevens, Effect of thrust coefficient on the flow blockage
waves, J. Fluid Mech. 862 (2019) 990–1028, http://dx.doi.org/10.1017/jfm. effects in closely-spaced spanwise-infinite turbine arrays, J. Phys. Conf. Ser. 1618
2018.969. (2020) 062069, http://dx.doi.org/10.1088/1742-6596/1618/6/062069.
[26] S.N. Gadde, R.J.A.M. Stevens, Effect of turbine-height on wind farm performance [49] R.J.A.M. Stevens, J. Graham, C. Meneveau, A concurrent precursor inflow
in the presence of a low-level jet, J. Renew. Sustain. Energy 13 (2021) 013305, method for large eddy simulations and applications to finite length wind farms,
http://dx.doi.org/10.1063/5.0026232. Renew. Energy 68 (2014) 46–50, http://dx.doi.org/10.1016/j.renene.2014.01.
[27] S.N. Gadde, R.J.A.M. Stevens, Interaction between low-level jets and wind farms 024.
in a stable atmospheric boundary layer, Phys. Rev. Fluids 6 (2021) 014603, [50] A.K. Blackadar, Boundary layer wind maxima and their significance for the
http://dx.doi.org/10.1103/PhysRevFluids.6.014603. growth of nocturnal inversions, Bull. Am. Meteorol. Soc. 38 (5) (1957) 283–290,
[28] J.D. Albertson, Large Eddy Simulation of Land-Atmosphere Interaction Ph.D. http://dx.doi.org/10.1175/1520-0477-38.5.283.
dissertation, University of California, 1996. [51] C. Vanderwel, M. Placidi, B. Ganapathisubramani, Wind resource assessment
[29] J.D. Albertson, M.B. Parlange, Surface length-scales and shear stress: implications in heterogeneous terrain, Phil. Trans. R. Soc. A 375 (2017) 20160109, http:
for land-atmosphere interaction over complex terrain, Water Resour. Res. 35 //dx.doi.org/10.1098/rsta.2016.0109.
(1999) 2121–2132, http://dx.doi.org/10.1029/1999WR900094. [52] R.B. Cal, J. Lebrón, L. Castillo, H.S. Kang, C. Meneveau, Experimental study of
[30] S.N. Gadde, R.J.A.M. Stevens, Effect of Coriolis force on a wind farm wake, the horizontally averaged flow structure in a model wind-turbine array boundary
J. Phys. Conf. Ser. 1256 (2019) 012026, http://dx.doi.org/10.1088/1742-6596/ layer, J. Renew. Sustain. Energy 2 (2010) 013106, http://dx.doi.org/10.1063/1.
1256/1/012026. 3289735.
[31] S.N. Gadde, A. Stieren, R.J.A.M. Stevens, Large-eddy simulations of stratified [53] R.J.A.M. Stevens, D.F. Gayme, C. Meneveau, Large eddy simulation studies of the
atmospheric boundary layers: Comparison of different subgrid models, Boundary- effects of alignment and wind farm length, J. Renew. Sustain. Energy 6 (2014)
Layer Meteorol. 178 (2021) 363–382, http://dx.doi.org/10.1007/s10546-020- 023105, http://dx.doi.org/10.1063/1.4869568.
00570-5.
[32] Á. Jiménez, A. Crespo, E. Migoya, Application of a LES technique to characterize
the wake deflection of a wind turbine in yaw, Wind Energy 13 (2010) 559–572,
http://dx.doi.org/10.1002/we.380.

58

You might also like