• No results found

Large eddy simulations of the effect of vertical staggering in large wind farms

N/A
N/A
Protected

Academic year: 2021

Share "Large eddy simulations of the effect of vertical staggering in large wind farms"

Copied!
16
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

R E S E A R C H A R T I C L E

Large eddy simulations of the effect of vertical staggering in

large wind farms

Mengqi Zhang

1,2

Mark G. Arendshorst

1

Richard J. A. M. Stevens

1

1Physics 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

2Department of Mechanical Engineering,

National University of Singapore, 10 Kent Ridge Crescent, 119260, Singapore

Correspondence

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, PO Box 217, 7500 AE Enschede, The Netherlands. Email: r.j.a.m.stevens@utwente.nl

Funding information

Stichting voor de Technische Wetenschappen (STW), Grant/Award Number: VIDI grant 14868 ; Fundamenteel Onderzoek der Materie (FOM), Grant/Award Number: CSER TENURE TRACK GRANT

Abstract

In order to study the effect of vertical staggering in large wind farms, large eddy simulations (LES) of large wind farms with a regular turbine layout aligned with the given wind direction were conducted. In the simulations, we varied the hub heights of consecutive downstream rows to create vertically staggered wind farms. We analysed the effect of streamwise and spanwise turbine spacing, the wind farm layout, the turbine rotor diameter, and hub height difference between consecutive downstream turbine rows on the average power output. We find that vertical staggering significantly increases the power production in the entrance region of large wind farms and is more effective when the streamwise turbine spacing and turbine diameter are smaller. Surprisingly, vertical staggering does not significantly improve the power production in the fully developed regime of the wind farm. The reason is that the downward vertical kinetic energy flux, which brings high velocity fluid from above the wind farm towards the hub height plane, does not increase due to vertical staggering. Thus, the shorter wind turbines are effectively sheltered from the atmospheric flow above the wind farm that supplies the energy, which limits the benefit of vertical staggering. In some cases, a vertically staggered wind farm even produced less power than the corresponding non vertically staggered reference wind farm. In such cases, the production of shorter turbines is significantly negatively impacted while the production of the taller turbine is only increased marginally.

KEYWORDS

atmospheric boundary layer, large eddy simulation, power production, turbulence, vertically staggered wind farm, wake model

1

INTRODUCTION

Wind power is a very promising clean and sustainable energy form. The recent decades have shown a rapid development of the wind industry with an ever increasing contribution to the total energy production worldwide.1A recent trend is the construction of very large wind farms. The flow dynamics in these wind farms is complex due to the interaction between the different wind turbine wakes and the atmospheric flow above the wind farm. Unfortunately, the operation of wind turbines is negatively impacted by wakes created by upstream turbines. For reviews on wind farm modeling, we refer to other studies.2-4An overview of research challenges for the wind energy community is discussed by van Kuik et al.1 In order to reduce the impact of wakes, and to increase the energy conversion efficiency of wind turbines, it is common to use horizontal staggering or more advanced configurations in large wind farms. Advanced wind farm layouts limit wake effects and ensure a good wind farm performance. An extensive review on layout optimization techniques is given by Herbert-Acero et al.5While horizontal variations in the wind farm layout have been studied extensively, the concept of using vertical staggering to limit wake effects has been relatively unexplored.

Most studies on vertically staggered wind farms focus on the use of simple analytical wake models to explore the potential of vertical staggering to optimize the wind farm layout using various optimization techniques. Chen et al6use a genetic algorithm in combination with the

. . . . This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium, provided the original work is properly cited.

© 2018 The Authors Wind Energy Published by John Wiley & Sons Ltd

(2)

Jensen model7to determine the optimal wind farm layout and find that vertical staggering can increase wind farm power output when the hub height difference between consecutive downstream rows is sufficiently large. Chen et al8employed a three-dimensional greedy algorithm to optimize both turbine positions and hub heights in terms of a predefined cost function per unit power for flat and complex terrain while using the Jensen and particle wake model9to account for wake effects. They find that using turbines of different heights allows one to harness more energy from the atmospheric winds at different elevations. DuPont et al10used the ‘‘extended pattern search multiagent system algorithm’’ to optimize wind farm power outputs evaluated by the PARK model7,11for stable, neutral, and unstable atmospheric conditions. They find that for a wind farm subjected to a dominant wind direction and stable atmospheric conditions, using turbines of intermediate height is optimal. However, for unstable atmospheric conditions and strongly varying incoming wind direction, the optimal wind farm consists of fewer but significantly larger turbines. These results show that the optimal configuration in a complex situation, involving different control parameters, can change dramatically. Vasel-Be-Hagh and Archer12conducted an optimization analysis of the annual energy production (AEP) of the Lillgrund farm in Sweden by varying the wind turbine hub heights. They employed a greedy search algorithm to find the optimal hub height between Hminand

Hmax(which are the predetermined minimum and maximum hub heights, respectively) for all turbines. Using the PARK model, they predict that vertical staggering can increase the AEP by2%. In the corresponding optimal configuration, all turbines have a height of either Hminor Hmax, ie, the benefit of vertical staggering is limited by the hub height range that is considered feasible.

The benefit of analytical wake models is that they are fast and easy to use. Such methods are crucial for wind farm layout optimizations in which the performance of thousands of wind farm layouts for all wind directions needs to be evaluated.5As discussed by Göçmen et al,3 analytical wind farm models provide very useful insight into wind farm performance for the community. To further improve analytical wake models, comparison to experimental and simulation data, which can capture the complex flow dynamics in very large wind farms4,13-16in more detail, is required. Considering the potential benefit of vertical staggering observed in analytical models, there is a need to verify whether these findings are supported by results from experiments or large eddy simulations (LES). LES have now developed so far that they can provide valuable information about wind farm performance.2,4,17However, experiments or LES of vertically staggered wind farms are relatively rare to date.

Chamorro et al18studied a vertically staggered wind farm concept in which turbines with different hub height and turbine diameter were used to increase the wind farm power output. They performed wind tunnel measurements in a scaled down wind farm of 3 by 8 turbines focusing on the effect of the surface roughness. They observe that, compared with the homogeneous-size wind farm, the vertically staggered wind farm with turbines of different size and height can generate a higher turbulence mixing in the shear layer. This results in a higher vertical kinetic energy flux that brings down high velocity wind from above the wind farm, which results in a higher potential energy harvesting. Vested et al19performed wind tunnel measurements on the effect of vertical staggering in wind farms with four downstream turbine rows. They observe an increase in the power production of up to25% due to vertical staggering. We note that in these experiments, the base height of all turbines is kept constant,

while half of the turbines is increased in height compared with the reference case to obtain vertical staggering. Thus, part of the observed power increase has to be attributed to the higher production potential of the considered vertically staggered wind farms.

Wang et al20compared analytical wake model solutions (PARK, Larsen,21and EPFL wake model22) with results obtained from the commercial software package FLUENT. They find that the wind turbine wakes behave differently in vertically staggered wind farms, and this influences the ability of wake models to capture this behavior. Another study by Vasel-Be-Hagh and Archer12compared results from wake model calculations with LES results obtained with SOWFA.23They find that the Lillgrund wind farm with optimized hub heights generates about6.3% more power in LES while, as mentioned above, the evaluations using the PARK model predict a2% increase.

Xie et al24 used LES to consider a different scenario, namely, the deployment of smaller vertical-axis wind turbines in an already existing wind farm of horizontal-axis wind turbines. They found that the addition of the vertical-axis wind turbines leads to an increase in the energy production, because the wakes recover faster due to the enhanced turbulence created by the vertical-axis wind turbines. These results were confirmed in an analytical top-down model they introduced.

Based on the literature review above, it is clear that most research on vertical staggering in wind farms has focused on the use of analytical wake models to investigate the potential benefit of vertical staggering. However, there is still limited reference data, which can be used to verify and improve model calculations, available from experiment or LES. In this work, we present data from LES in which we study the effect of streamwise and spanwise turbine spacing, turbine rotor diameter, the hub height difference between consecutive turbine rows, and the wind farm layout, on the power output of vertically staggered wind farms. We start with a description of the numerical methods in section 2. In section 3, we present an analysis of the power output of vertically staggered wind farms, the vertical profiles of the streamwise velocity and vertical kinetic energy flux, and a comparison of the LES results with Jensen model predictions. We conclude the paper with a discussion in section 4, while the details about the Jensen model are presented in Appendix A.

2

PROBLEM FORMULATION AND NUMERICAL METHODS

A sketch of the vertically staggered wind farm configuration and the corresponding design parameters is shown in Figure 1. The dimensionless streamwise turbine spacing sxand the spanwise turbine spacing sy(not shown in the sketch) are measured in terms of the turbine diameter D

and indicate the distance between two adjacent turbines in either direction. Panel A shows that in the reference aligned wind farm, all turbines have the same hub height zhand that all downstream turbines are entirely immersed in the wake of upstream turbines. Consequently, the power

(3)

(B)

(A)

FIGURE 1 Side view of the conceptual configuration of A, the aligned reference and B, the vertically staggered wind farm layout. The streamwise direction is along positive x, and the gray areas represent linearly expanding wakes. sxis the dimensionless streamwise turbine spacing. The

average turbine hub height is zh. Hdmeasures the height difference with respect to zh, so the height difference between two consecutive turbine

rows is2Hd. In this study, the even turbine rows have tall turbines and the odd rows short turbines, unless mentioned otherwise

production of all downstream wind turbines is much lower than for turbines in the first row. Placing downstream turbines such that they are (partially) outside the wake of upstream turbines is one of the possibilities to enhance the overall power production of a wind farm. This is the idea behind the vertically staggered wind farm shown in panel B. The figure reveals that part of the second wind turbine is outside the wake of the first turbine when the first turbine is lowered, and the second one is elevated. The degree of vertical staggering is dictated by the elevation/abasement

Hdrelative to the averaged hub height zh, ie, the height difference between two consecutive turbine rows is2Hd. We use wind farms with an

even number of rows, lowering the odd rows and increasing the height of the even rows, so the average hub height remains the same. This ensures that the overall power generating potential, ie, assuming there are no wake effects, of all considered wind farms is roughly the same. In fact, due to the logarithmic velocity profile in the atmospheric boundary layer (ABL), the potential power generation of vertically staggered wind farm can be about3% to4% lower than for the aligned reference wind farm, see section 3.3 for details. In this work, we use LES to simulate the

turbulent flow in very large vertically staggered wind farms and evaluate their performance compared with the aligned reference case. In the fully developed regime of the wind farm (effectively after the fourth row), the wind farm performance does not depend on the ordering of the short and tall turbines. However, as explained at the end of section 3.1, a larger benefit of vertical staggering can be obtained in the entrance region of the wind farm (especially the first two rows) when the odd rows (and thus the first row) would have been elevated instead of the even rows.

2.1

Large eddy simulation

The LES code we use solves the incompressible Navier-Stokes equations and the continuity equation

𝜕t̃ui+𝜕j(̃uĩuj) = −𝜕ĩp∗−𝜕j𝜏ij𝛿i1𝜕1p∞∕𝜌 + fi, (1)

𝜕ĩui= 0, (2)

where ̃ui represents the filtered velocity field, ̃p∗ = ̃p∕𝜌 + 𝜏kk∕3 − p∞∕𝜌is the filtered modified pressure,𝜌is the fluid density, and𝜏ijis the

subgrid-scale stress. The turbine forcing is represented by a body force fi. Coriolis and thermal effects are not specifically included, an approach

which was also used in previous studies.25-28 The code employs a pseudospectral method in horizontal directions and a second order finite difference scheme in the vertical direction. Time integration is performed using a second-order Adams-Bashforth scheme. The subgrid scale dynamics are modeled by the Lagrangian-averaged scale-dependent dynamic model.29The wall stress at the ground is captured by a standard rough wall model30using velocities test filtered twice at the grid scale.29For the top boundary, we use a zero vertical velocity and zero shear stress boundary condition. A detailed description of the basic flow solver can be found in Albertson and Parlange.31In order to consider finite size wind farms, we use the concurrent precursor inflow method introduced in Stevens et al.32Each time step, a part of the flow field from this simulation is used to provide the inflow condition for a second section in which the wind farm is placed. In the wind farm section, which is periodic due to the use of spectral methods in the horizontal direction, a long fringe region at the end of the computational section is used to ensure a smooth transition from the flow formed behind the wind farm towards the applied inflow condition. To reduce the effect of the location of the high and low velocity speed streaks in the ABL, we perform the simulations for a long time and very slowly shift the entire flow in the inflow generating domain in the spanwise direction to get well converged (streak independent) results. This method is tested in Stevens et al,33and the benefits of such a method are discussed in more detail by Munters et al.34Our LES code has been validated against EPFL wind tunnel experiments on a10 × 3wind turbine array,33,35,36the Horns Rev wind farm measurements,32and high-resolution wake measurements performed at Delft.37 We use a computational domain of4𝜋H×𝜋H×H in streamwise, spanwise, and vertical directions. Here, the vertical height of the domain H = 1000m and the roughness height at the ground is z0,lo = 0.1m, which corresponds to open terrain according to the Davenport classification of effective roughness.38,39The corresponding streamwise turbulence intensity at 100 m is12.2%. We use a grid of512 × 128 × 241nodes, which is sufficient to capture the large-scale dynamics in the wind farm and similar to the grid resolution in previous studies, see for example other studies.26,32,40The turbines are modeled using an actuator disk model,25,32,41which has been shown to reasonably accurately capture wake profiles further downstream with a much lower computational cost25,32,33,41,42than an actuator line model.43,44Following previous studies,25 we assume that the average turbine power output is equal to the mechanical energy loss in the fluid, ie,P = −FUd, where the overbar indicates

(4)

TABLE 1 Wind farm configurations simulated in this studya

Case Name zh, m D, m Hd, m Number of Turbines (Nx× Ny) sx sy Figure

a0 − a4 100 100 [0;10;20;30;40] 18 × 6 5.24 5.24 2 to 5, 7 to 9, 11 b0 − b4 100 100 [0;10;20;30;40] 14 × 6 6.98 5.24 3, 4, 11 c0 − c4 100 100 [0;10;20;30;40] 8 × 6 10.47 5.24 3, 4, 11 A0 − A3 120 150 [0;10;20;30] 12 × 4 5.24 5.24 3, 4, 11 B0 − B3 120 150 [0;10;20;30] 8 × 4 6.98 5.24 3, 4, 11 C0 − C3 120 150 [0;10;20;30] 6 × 4 10.47 5.24 3, 4, 11 d0∕d4 100 100 [0;40] 18 × 6 5.24 2.62 5 e0∕e4 100 100 [0;40] 18 × 6 5.24 7.85 5 f0∕f4 100 100 [0;40] 18 × 6 5.24 5.24 6 F0∕F4 120 150 [0;30] 12 × 4 5.24 5.24 6 g0∕g4 100 100 [0;40] 18 × 6 5.24 5.24 7

aThe cases with smaller (D

s = 100 m and zh = 100 m) and larger (Dl= 150 m and zh = 120 m) turbines are denoted by

normal and capital letters, respectively. The number in the case name refers to the height difference Hd, and Nxand Ny

indicate the number of turbines in streamwise and spanwise direction, respectively. sxand syindicate the dimensionless

turbine spacing in streamwise and spanwise direction. For the cases with small wind turbines, we normalize the turbine spacings using Dsand for the cases with larger turbines using Dl. In cases a to e (and g), the ordering of tall and short

turbines is as indicated in Figure 1, while in case f, the tall turbines are located in the even rows and the short turbines in the odd rows. In cases a to f, a ‘‘horizontally’’ aligned configuration is considered, while in case g, a ‘‘horizontally’’ staggered wind farm is simulated.

time averaging. Here,F = −12C

T𝜌U

2

dAis the force used in the actuator disk model with Udthe disk averaged velocity, A = 𝜋D

2

∕4is the turbine rotor area, and CT= CT∕(1 −a)2, where a is the axial induction factor. We use a constant thrust coefficient C

T = 3∕4, which is representative

for turbines in utility scale wind farms.

We consider two kinds of wind turbines, smaller turbines defined as those with diameter Ds = 100m and hub height zh = 100m and larger

turbines defined as those with a diameter Dl = 150m and hub height zh = 120m. The larger wind turbine is considered because of the recent

development trend towards such turbines in industry.1We keep the simulation area constant in all simulations. In the main set of simulations, see cases a to c in Table 1; we keep the dimensionless spanwise spacing (sy = 5.24) constant, while we vary the dimensionless streamwise spacing sx. For the wind farms with the smaller turbines, the number of turbines in streamwise and spanwise directions varies from18 × 6(sx = 5.24) to

14 × 6(sx = 6.98) to9 × 6(sx = 10.47). For the simulations with larger turbines, the number of turbines ranges from12 × 4(sx= 5.24) to9 × 4

(sx = 6.98) to6 × 4(sx = 10.47). For the smaller turbines, we consider Hd = 0, 10, 20, 30, 40m, and for the larger turbines Hd = 0, 10, 20, 30m,

see Figure 1 for a definition of Hd. We limited the considered Hdrange by requiring that the bottom of the turbine is at least 10 m from the

ground. This puts the maximum turbine hub height considered in this study at 150 m, which is ambitious, but realistic, considering wind turbine development trends. Cases d and e, see Table 1, in combination with case a allow us to investigate the influence of the spanwise turbine spacing on the effect of vertical staggering for a fixed streamwise turbine spacing sx = 5.24. In case f, we use the same wind farm layout as in case a but

locate the tall turbines in the odd rows and short turbines in the even rows, instead of placing the tall turbines in the even rows, to asses the effect of the vertical staggering order in the entrance region of the wind farm.45In case g, see Table 1, we simulate a ‘‘horizontally’’ staggered wind farm with sx =sy = 5.24in order to asses the effect of the wind farm layout.

Due to the computational requirements for LES, a number of simplifying assumptions have to be made. This is also the reason that the development of computationally less demanding modeling methods is very important. In the present work, we restrict ourselves to neutral atmospheric thermal stability conditions and do not include the effect of the Coriolis force. As is explained in the appendix of Calaf et al,25this is equivalent to assuming a given value for the transverse geostrophic wind, and this approach is expected to give meaningful results as the flow in the inner region (up to0.15 − 0.20H) is not much influenced by external effects such as Coriolis forces. In addition, we assume that all turbines operate in regime II,46in which the thrust coefficient C

Tis assumed to be constant, independent of wind speed, and we neglect the variation

of the power coefficient CPwith wind speed. These are reasonable assumptions since LES47and field measurement13generally report data for

narrow wind speed bins, typically≈ 8 ± 0.5m/s, which corresponds to turbines operating in regime II. According to Porté-Agel et al,47this regime II is particularly interesting as turbines operate most of the time under these conditions.

3

RESULTS

Figure 2A shows the typical wake pattern observed in the reference aligned wind farm, while Figure 2B reveals that in the vertically staggered wind farm, the taller turbines extrude further into the ABL. Clearly, the flow is much more complex in the vertically staggered wind farm than in the aligned reference case. In section 3.1, we first evaluate the power output of vertically staggered wind farms compared with the corresponding non vertically staggered reference wind farm. Subsequently, we will analyse the streamwise velocity and vertical kinetic energy flux profiles in section 3.2. Finally, the LES results are compared with the Jensen model predictions in section 3.3.

(5)

FIGURE 2 Vertical cross section of the spanwise and time-averaged streamwise velocity⟨u⟩normalized with the friction velocity u∗for A, the aligned reference wind farm a0and B, the vertically staggered wind farm a4. The black vertical lines indicate the turbine locations [Colour figure can be viewed at wileyonlinelibrary.com]

(A)

(B)

(C)

(D)

(E)

(F)

FIGURE 3 Normalized power output PPref,1as function of downstream position. Pref,1is the power output of the first turbine row in the reference aligned wind farm. Left column: smaller wind turbines with Ds = 100m and zh = 100m; right column: larger wind turbines with Dl = 150m and zh = 120m. Each row shows the result for a different streamwise turbine spacing sx. The turbine distances are made

(6)

(A)

(B)

(C)

(D)

(E)

(F)

FIGURE 4 Normalized cumulative power output as function of the downstream position, see details in section 3.1. Left column: smaller wind turbines with Ds = 100m and zh = 100m; right column: larger wind turbines with Dl = 150m and zh = 120m. Each row shows the result for a

different streamwise turbine spacing sx. The turbine distances are made dimensionless using either Dsor Dl. The insets in panel A and B provide

a zoom in for the results in the fully developed regime of the wind farm [Colour figure can be viewed at wileyonlinelibrary.com]

3.1

Power output

Figure 3 shows the power production as a function of downstream direction for different vertically staggered wind farms. The results are normalized with the power production of the first row of the corresponding reference aligned wind farm, which is denoted as Pref,1. The results are averaged over time and in spanwise direction over six (smaller turbines) or four (larger turbines). The statistical uncertainty in the presented data is about2%. This number reflects the difference in the power production obtained in the first and second half of the considered simulation

interval. The spin-up time required to reach the statistical stationary state is excluded from all data analysis. Figure 3 shows that the power output of the second row drops significantly compared with the production of the first row due to wake effects. Further downstream, the power output of turbines increases slightly due to the creation of a strong downward vertical kinetic energy flux that helps the wakes to recover faster.25,48-51 Depending on sx, the production in the fully developed regime asymptotes to around 50% to 70% of the production in the first row. These results

are in agreement with results obtained in previous studies.50,52

Figure 3 reveals that the power output of the first row decreases with increasing Hd, while the production of the second row increases with

increasing Hd. This result is expected since in an ABL, with a logarithmic mean velocity profile, the shorter wind turbines are exposed to a weaker

incoming velocity, and their production is accordingly lower; vice versa, the taller wind turbines can produce more power as these turbines can tap into the higher wind speed regions. In addition, the turbines on the second row are displaced more outside the wake of the preceding turbines when Hdis increased. The combined production of the first two turbine rows can be significantly higher than for the reference aligned wind farm,

see also Figure 4. So vertical staggering can strongly increase the power output in the entrance region of a wind farm.

Figure 3 shows that in the fully developed regime of the wind farm, the power production is higher for taller turbines (even rows) and lower for shorter turbines (odd rows). The reason for this zigzag pattern is the same as discussed before, namely, that the taller turbines are partially outside

(7)

the wake of the preceding shorter turbines due to which the taller turbines have better access to the undisturbed atmospheric flow above the wind farm. Figure 3 shows that these trends are similar for the wind farms with smaller (Ds = 100m and zh = 100m) and larger (Dl = 150m and zh = 120m) turbines. One difference, however, is that in wind farms with larger turbines, the second turbine row always produces less power than the first row, while for the wind farms with smaller turbines, the second row can produce more power than the first row. The reason is that it is much easier to place the turbines on the second row significantly outside the wakes created by the turbines in the first row when smaller turbines are used.

The benefit of vertical staggering can be judged better by plotting the normalized cumulative power output, which compares to power output of the vertically staggered wind farms with the aligned reference wind farm. We determine the normalized cumulative power production as a function of the downstream position by comparing the power production of the first x rows of the vertically staggered wind farm with the power production of the first x rows of the aligned reference wind farm. This normalized cumulative production, denoted as PPref,c, thus indicates whether vertical staggering increases or decreases the production compared with the reference wind farm. The normalized cumulative production for the reference wind farm is one by construction. Figure 4 shows, in agreement with Figure 3, that vertical staggering can be highly beneficial in the entrance region of the wind farm as in panel A, a benefit up to20% is observed. The figure also reveals that the benefit of vertical staggering

is smaller for larger turbines and decreases with increasing streamwise turbine spacing and decreasing Hd.

Figure 4 shows that vertical staggering seems only marginally beneficial in the fully developed regime of the wind farm. All graphs namely show that the normalized cumulative power production decreases with increasing downstream distance for all vertically staggered wind farm configurations considered here and only for the wind farms with a small dimensionless streamwise spacing a benefit of about5% remains at the

end of the wind farm. Note that Figure 11 shows that the power production in the last four rows of the vertically staggered wind farms is slightly lower than in the last four rows of the corresponding aligned reference wind farm for most cases. A closer investigation of Figure 3 reveals that

(A)

(B)

(C)

(D)

(E)

(F)

FIGURE 5 Effect of vertically staggering in wind farms with small turbines (Ds = 100m and zh = 100m) and a spanwise spacing syof A, sy = 2.62; B, sy = 5.24; and C, sy = 7.85. The left panels show the normalized power output PPref,1as function of downstream position. Pref,1 is the power output of the first turbine row in the reference aligned wind farm. B, The right panels show the normalized cumulative power production, see details in section 3.1 [Colour figure can be viewed at wileyonlinelibrary.com]

(8)

FIGURE 6 Normalized power output PPref,1as function of downstream position. Pref,1is the power output of the first turbine row in the reference aligned wind farm for A, smaller (Ds = 100m and zh = 100m) and B, larger (Dl = 150m and zh = 120m) turbines. For the wind farm

with smaller turbines, the performance increase with respect to the aligned reference wind farm for the first four turbine rows is15% when the

odd turbine rows are elevated and12% for when the odd turbine rows are lowered. For the wind farm with the larger turbines, these numbers

are10% and4%, respectively. For more details, we refer to Zhang et al45[Colour figure can be viewed at wileyonlinelibrary.com]

FIGURE 7 Effect of vertical staggering in a horizontally ‘‘staggered’’ wind farm with smaller turbines (Ds = 100m and zh = 100m), and sx = 5.24and sy = 5.24. A, Normalized power output PPref,1as function of downstream position. Pref,1is the power output of the first turbine row in the reference aligned wind farm. B, The normalized cumulative power production, see details in section 3.1 [Colour figure can be viewed at wileyonlinelibrary.com]

this effect is mainly caused by the poor performance of the shorter turbines in the fully developed regime of the wind farm, which presumably is caused by the slow wake recovery near the ground.

Figure 5 shows that vertical staggering is slightly more beneficial for a wind farm with a small spanwise spacing of sy = 2.62than for wind

farms with larger spanwise turbine spacing (sy = 5.24and sy = 7.85) for which sideways wake effects between different turbine columns are

negligible. Figure 6 shows that a vertically staggered wind farm in which the odd (and therefore the first) turbine rows are elevated produces a higher power output in the entrance region of the wind farm than a wind farm in which the even (and therefore the second) rows are elevated. The reason for this is that the first turbine row always produces a lot more power than the second row. Therefore, lowering the first turbine row leads to a drastic drop of the most productive turbines in the entire farm (tapping into lower wind speeds from the log profile), which is difficult to compensate for with a taller second turbine (and other even turbines) because of the vertically expanding wakes. Further downstream, in the fully developed regime of the wind farm, the performance is identical for both configurations and does not depend on the order of the tall and short turbines. As we focus on the benefit of vertical staggering in large wind farm, in particular on the performance in the fully developed regime, our initial choice to consider vertically staggered wind farms in which the even rows are elevated does not influence the discussion in the remainder of the manuscript. However, we emphasize that in the entrance region of the wind farm, larger benefits of vertical staggering can be obtained than reported here by elevating the first instead of the second turbine row. For a more detailed discussion on this, we refer to Zhang et al.45

As wake effects are strongest for the aligned configuration, the potential to improve the power production using vertical staggering is largest for the ‘‘horizontally’’ aligned wind farms. So far, we have seen that even for this ‘‘best case’’ scenario vertical staggering does not lead to a significant increase in the power production in the fully developed regime of the wind farm for most cases. In Figure 7, we show that in a ‘‘horizontally’’ staggered wind farm, the use of vertical staggering can even reduce the power production in the fully developed regime of the wind farm compared with the corresponding non vertically staggered wind farm. In the remainder of the manuscript, we focus our analysis on cases a to c, see Table 1, to get a better understanding of the reason why vertical staggering does not increase the power production in the fully developed regime of very large wind farms.

(9)

FIGURE 8 Horizontally averaged streamwise velocity profiles as function of height for the wind farms a0to a4. The velocity is extracted1Ds

downstream of the mentioned row number. The gray dashed lines represent undisturbed ABL inflow profile. The vertical dashed lines indicate the vertical extend of the turbines in the vertically staggered wind farms [Colour figure can be viewed at wileyonlinelibrary.com]

3.2

Averaged velocity and vertical kinetic energy flux

In order to confirm the above conjecture that the poor performance of the shorter turbines in the fully developed regime is caused by the slow wake recovery near the ground, we turn our attention to the averaged streamwise velocity and vertical kinetic energy flux profiles to study the effect of vertical staggering in more detail. Figure 8 reveals that at z = 100m, the maximum velocity deficit compared with the incoming flow decreases with increasing Hd. However, closer to the ground, the velocity deficit increases with Hd. The reason is that in a vertically staggered

wind farm, there are now also turbines closer to the ground. In fact, the figure shows that especially behind the short turbines (odd rows, right column of the figure), the velocity close to the ground is very low, and based on the velocity profiles measured behind the taller turbines, it is clear that the flow close to the ground only recovers slowly.

It was shown by Cal et al48 and Calaf et al25 thatΦ = −uuw, where the overbar indicates time averaging, is the dominant term of the turbulence-induced vertical flux of mean kinetic energy in large wind farms. It has been shown in several studies47,50,51,53,54 that there is a significant correlation between this flux and the wind turbine power extraction, while Allaerts and Meyers55,56showed that in stable boundary layers, the streamwise energy flux is an additional quantity of interest and is important in the entrance region of the extended wind farms. As we only consider neutral ABL and focus on the performance in the fully developed wind farm region, the vertical kinetic energy flux is a more suitable quantity to consider. In fact, as we will show below, the analysis of the vertical kinetic energy flux clarifies that vertical staggering does not significantly improve power production in the fully developed wind farm region.

In Figure 9, we analyse the spanwise average of the local vertical energy energy flux⟨Φ⟩(z) = −⟨uuw′⟩, where⟨.⟩indicates spanwise averaging. Here, we note that evaluating the vertical flux as⟨Φ⟩(z) = −⟨uuwgives nearly the same result as using⟨Φ⟩(z) = ⟨uuw⟩ + ⟨vvw⟩ + ⟨www as the last two terms are very small compared with the dominant term. Figure 9A and 9B reveal that the vertical kinetic energy flux develops ‘‘similarly’’ as function of downstream position in vertically staggered wind farms as in the aligned reference wind farm.57This result is consistent with the observation that the power production in the fully developed regime of the wind farm does not change much due to vertical staggering. A more comprehensive analysis of the development of the vertical kinetic energy flux as function of downstream direction is given in Figure 9C and 9D. These figures show that the development of the vertical kinetic energy flux as function of downstream position is almost the same in the vertically staggered wind farms as in the aligned reference case. We observe that at z = 200m, the downward vertical kinetic energy increases at 1 km downstream, while at z = 300m, this increase becomes visible at 2 km downstream. This is related to the development of the internal boundary layer, as discussed by Chamorro et al36and in Stevens and Meneveau.4At 200 m, the vertical kinetic energy flux⟨Φ⟩reaches a plateau around x = 5km (or 10 turbine rows), while such a plateau is less clear at 300 m. We find that at z = 200m, the vertical kinetic energy flux shows a sawtooth pattern as function of the downstream location for larger Hd. Specifically, for case a4, we see that at this height,⟨Φ⟩is lower

at the location of the taller turbines (even rows) than at the location of the shorter turbines (odd rows). This result is explained in Figure 10, where we sketch the vertically staggered wind farm and the shear layers generated at the top of the wind turbines. Due to the convection in the

(10)

FIGURE 9 Spanwise averaged vertical kinetic energy flux⟨Φ⟩profile as function of height for A, the aligned reference(a0)and the vertically staggered(a3)wind farm. The flux is calculated1Dsdownstream of the respective turbine rows, and the gray dashed lines represent the profile for the incoming flow. The vertical dashed lines indicate the vertical extend of the turbines in the respective wind farms. The dashed magenta line denotes the thickness of the internal boundary layer, and its value is 850 m. The development of⟨Φ⟩at C (z= 200m) and D (z= 300m) as function of downstream location. Dashed and solid vertical lines indicate the position of the shorter (odd rows) and taller (even rows) turbines, respectively [Colour figure can be viewed at wileyonlinelibrary.com]

FIGURE 10 A side view sketch of vertically staggered wind turbines. The gray patterns represent linearly expanding wakes behind the turbines as in Figure 1. The turbulent shear layers at the top of turbines are illustrated with symbolic eddies of different sizes (the sizes of the vortices in the figure are not scaled according to the turbine size). Dashed vertical lines represent the position of shorter turbines (odd rows), and solid lines represent the position of the taller turbine (even rows) as in Figure 9. The horizontal lines indicate the vertical positions z = 200m and 300 m, corresponding to the results shown in Figure 9C,D

ABL, a shear layer develops and thickens downstream of the turbines, while the size of the vortices inside becomes larger. Therefore, it can be understood that at z = 200m, the vertical kinetic energy flux at position A should be higher than at position B as the shear layer thickens and the turbulence becomes weaker downstream due to the entrainment of the ABL flow. Clearly, this effect is stronger when Hdis larger, because

the wind and shear layer are stronger at higher elevations. Therefore, the manifestation of the sawtooth pattern in the vertical kinetic energy flux along streamwise direction is most pronounced for Hd = 40m, see Figure 9. At z = 300m, the sawtooth pattern is barely visible, because it

(11)

FIGURE 11 Power output of the vertically staggered (VS) wind farms normalized by the power output of the corresponding aligned reference wind farm. The LES results are indicated by symbols and the Jensen model predictions by the lines. The top panels show the results for the wind farms with smaller turbines (Ds = 100m and zh = 100m), and the lower panels show the results for wind farms with larger turbines

(Dl = 150m and zh = 120m). The left column shows the results for the first two rows, and the right column shows the results for the last four

rows [Colour figure can be viewed at wileyonlinelibrary.com]

3.3

Power prediction by Jensen model

As mentioned in section 1, there are already several studies that have used analytical wake models to optimize the hub height distribution in wind farms. In this section, we compare our LES results with predictions from the Jensen model, which is the most commonly used wake model in such wind farm layout optimization studies.5Based on the results discussed above, we can identify two dynamically different regimes, namely, the entrance region of the wind farm (here defined as the first two rows) and the fully developed regime of the wind farm (here defined as the last four rows). Figure 11 compares the power output in the entrance region of the vertically staggered wind farms with the power output obtained in the aligned reference wind farm. Figure 11A and 11C shows that vertical staggering is beneficial in the entrance region of the wind farm. The benefit of vertical staggering is larger for the smaller turbines (Ds = 100m and zh = 100m) than for the larger turbines (Dl = 150m

and zh = 120m) and increases with increasing Hdand decreasing streamwise turbine spacing. These performance trends in the entrance region

of the wind farm are quite well captured by the Jensen model, which slightly underestimates the benefit of vertical staggering here. For details about the used Jensen model, we refer to the Appendix A.

Figure 11B and 11D shows that in the fully developed regime of the wind farm, vertical staggering is hardly beneficial. For most cases, the production in the fully developed regime is even slightly lower than in the corresponding reference wind farm. The reason for this is twofold; first of all, the power production potential of the vertically staggered wind farm is slightly lower than for the reference wind farm due to the logarithmic velocity profile in the ABL⟨u(z)⟩∕u∗= 1∕𝜅 ln (z∕z0), where𝜅 = 0.4is the von Kármán constant. The magnitude of this effect is given by the large aspect ratio limit of the Jensen model results shown in Figure 11. In addition, as discussed above, the wind turbine wakes near the ground recover very slowly, due to which the production of the shorter turbines is negatively affected. In contrast to the LES results, which show that vertical staggering does not improve the production in the fully developed regime of the wind farm, the Jensen model predictions show a significant improvement of the power production in the fully developed regime of the wind farm, especially for the smaller turbines (Ds = 100m and zh = 100m) and small streamwise turbine spacings. We believe the Jensen model has difficulties to predict the performance in

the fully developed regime because the complex interaction between the wind farm dynamics and the atmospheric flow above is not specifically accounted for in the Jensen model.13-16To model such large scale interactions in more detail, we have previously developed the coupled wake boundary layer (CWBL),15,58,59which relies on a combination of the Jensen and a top-down model. However, the CWBL model cannot be applied here as the top-down model developed by Calaf et al25assumes that all turbines have the same hub height, while an extension of this top-model

(12)

presented by Xie et al24is only valid when there is a clear vertical separation between the shorter and taller turbines, and these models are not applicable for the present cases.

4

DISCUSSION AND CONCLUSIONS

In this work, we studied the effect of vertical staggering on the power production in large wind farms using high-fidelity numerical simulations. We systematically varied key parameters such as the streamwise turbine spacing sx, the turbine rotor diameter D, and the hub height difference

between consecutive turbine rows, see Figure 1. We find that vertical staggering can increase the power production in the entrance region of the wind farm (first two rows) by up to20%, see Figure 4A, compared with the corresponding aligned reference wind farm. It turns out that

a significant hub height difference between consecutive rows is required to obtain a significant benefit of vertical staggering, which is easier to achieve for smaller turbines (Ds = 100m and zh = 100m) than for larger turbines (Dl = 150m and zh = 120m) when a realistic hub

height range is considered. In addition, we find that the benefit of vertical staggering in the entrance region of the wind farm decreases with increasing streamwise turbine spacing sx. Thus, for smaller sx, a larger benefit of vertical staggering is expected. However, turbine loads would

be exceptionally high when the streamwise spacing is too small, and therefore, we did not consider streamwise spacings below sx = 5.24. In

contrast to the streamwise spacing, the spanwise spacing has a marginal effect on the effectiveness of using vertical staggering, but we find that the power production in the entrance region of the wind farm can be increased more when the odd rows (and thus the first row) are elevated instead of the even rows, which was done for most cases in this study.

Surprisingly, we find that vertical staggering is much less beneficial in the fully developed regime of the wind farm than originally anticipated. In fact, the LES results show that vertical staggering hardly improves the production in the fully developed regime compared with the aligned reference wind farm. The reason for this result is twofold, ie, the power production potential (assuming there are no wake effects) of the vertically staggered wind farm is slightly lower (up to about3% to4%) than that of the corresponding aligned reference wind farm due to the

logarithmic velocity profile in the ABL. In addition, the vertical kinetic energy flux, which brings high velocity wind from above the farm down to the hub height plane, does not increase due to vertical staggering. As a result, the wakes close to the ground recover slowly, and as a result, the performance of the shorter turbines in the fully developed regime can be severely impacted. We also find that the prevailing wind direction has a significant effect on the benefit of vertical staggering. In a ‘‘horizontally’’ staggered wind farm with sx = sy = 5.24, we namely find that vertical

staggering reduces the power output compared with the corresponding reference case as a result of the negative aspects of vertical staggering discussed above. Based on these results, we do not expect to find wind directions for which the considered type of vertical staggering will greatly enhance the power production in the fully developed regime of the wind farm.

Obviously, the conclusions reached here are still subject to considerable limitations as the LES have been performed for idealized conditions, such as neutral conditions and assume a flat terrain with no topography effects. Clearly, the effects of vertical staggering can be very different when such conditions are also taken into account. In addition, we emphasize that we only considered one class of vertically staggered wind farms, assuming for example that all wind turbines have the same size. As discussed in section 1, some different vertical staggering concepts have been proposed such as the use of turbines with different size and height18or a combination of horizontal- and vertical-axis wind turbines.24Different relative turbine sizes can greatly influence the possibility to use vertical staggering to place turbines outside the wakes of proceeding turbines and the corresponding wind farm dynamics. Further work on these aspects is required to get a better insight into the possibilities to use vertical staggering effectively. In addition, the development of simple analytical models that can accurately capture the benefit of vertical staggering in the fully developed regime of the wind farm will be very interesting.

ACKNOWLEDGEMENTS

We gratefully acknowledge the insightful comments and suggestions by anonymous referees during the review process. This work is part of the Shell-NWO/FOM-initiative Computational sciences for energy research of Shell and Chemical Sciences, Earth and Live Sciences, Physical Sciences, FOM, and STW and an STW VIDI grant (No. 14868). This work was carried out on the national e-infrastructure of SURFsara, a subsidiary of SURF cooperation, the collaborative ICT organization for Dutch education and research.

ORCID

Richard J. A. M. Stevens http://orcid.org/0000-0001-6976-5704

REFERENCES

1. van Kuik GAM, Peinke J, Nijssen R, et al. Long-term research challenges in wind energy—a research agenda by the European Academy of Wind Energy. Wind Energ Sci. 2016;1:1-39.

2. Breton SP, Sumner J, Sørensen JN, Hansen KS, Sarmast S, Ivanell S. A survey of modelling methods for high-fidelity wind farm simulations using large eddy simulation. Philos Trans A Math Phys Eng Sci. 2017;375:20160097.

3. Göçmen T, van der Laan P, Réthoré P-E, Diaz AP, Larsen GC, Ott S. Wind turbine wake models developed at the technical university of Denmark: a review. Renew Sust Energ Rev. 2016;60:752-769.

(13)

4. Stevens RJAM, Meneveau C. Flow structure and turbulence in wind farms. Annu Rev Fluid Mech. 2017;49:311-339. http://dx.doi.org/10.1146/ annurev-fluid-010816-060206

5. Herbert-Acero JF, Probst O, Réthoré PE, Larsen GC, Castillo-Villar KK. A review of methodological approaches for the design and optimization of wind farms. Energies. 2014;7(11):6930-7016. https://doi.org/10.3390/en7116930

6. Chen Y, Li H, Jin K, Song Q. Wind farm layout optimization using genetic algorithm with different hub height wind turbines. Energy Convers Manag. 2013;70:56-65. https://doi.org/10.1016/j.enconman.2013.02.007

7. Jensen NO. A note on wind generator interaction. Risø-M-2411, Roskilde, Risø National Laboratory; 1983.

8. Chen K, Song MX, Zhang X, Wang SF. Wind turbine layout optimization with multiple hub height wind turbines using greedy algorithm. Renew Energy. 2016;96, Part A:676-686. https://doi.org/10.1016/j.renene.2016.05.018

9. Song MX, Chen K, He ZY, Zhang X. Wake flow model of wind turbine using particle simulation. Renew Energy. 2012;41:185-190. https://doi.org/10. 1016/j.renene.2011.10.016

10. DuPont B, Cagan J, Moriarty P. An advanced modeling system for optimization of wind farm layout and wind turbine sizing using a multi-level extended pattern search algorithm. Energy. 2016;106:802-814. https://doi.org/10.1016/j.energy.2015.12.033

11. Katic I, Højstrup J, Jensen NO. A Simple Model for Cluster Efficiency. In: W Palz & Sesto E eds. EWEC'86. Proceedings. Vol. 1 Rome: A. Raguzzi; 1987:407-410.

12. Vasel-Be-Hagh A, Archer CL. Wind farm hub height optimization. Appl Energy. 2017;195:905-921. https://doi.org/10.1016/j.apenergy.2017.03.089 13. Barthelmie RJ, Hansen K, Frandsen ST, et al. Modelling and measuring flow and wind turbine wakes in large wind farms offshore. Wind Energy.

2009;12:431-444.

14. Son E, Lee S, Hwang B, Lee S. Characteristics of turbine spacing in a wind farm using an optimal design process. Renew Energy. 2014;65:245-249. 15. Stevens RJAM, Gayme DF, Meneveau C. Coupled wake boundary layer model of wind-farms. J Renew Sustainable Energy. 2015;7:23115. http://dx.

doi.org/10.1063/1.4915287

16. Wu YT, Porté-Agel F. Modeling turbine wakes and power losses within a wind farm using LES: an application to the Horns Rev offshore wind farm.

Renew Energy. 2015;75:945-955.

17. Mehta D, van Zuijlen AH, Koren B, Holierhoek JG, Bijl H. Large eddy simulation of wind farm aerodynamics: a review. J Wind Eng Ind Aerodyn. 2014;133:1-17.

18. Chamorro LP, Tobin N, Arndt REA, Sotiropoulos F. Variable-sized wind turbines are a possibility for wind farm optimization. Wind Energy. 2014;17:1483-1494. https://doi.org/10.1002/we.1646

19. Vested MH, Hamilton N, Sørensen JN, Cal RB. Wake interaction and power production of variable height model wind farms. J Phys Conf Ser. 2014;524:12169.

20. Wang L, Tan AC, Cholette M, Gu Y. Comparison of the effectiveness of analytical wake models for wind farm with constant and variable hub heights.

Energy Conversion and Management. 2016;124:189-202. https://doi.org/10.1016/j.enconman.2016.07.017

21. Larsen GC. Simple wake calculation procedure. Risø-M-2760, Roskilde, Risø National Laboratory; 1988.

22. Bastankhah M, Porté-Agel F. A new analytical model for wind-turbine wakes. Renew Energy. 2014;70:116-123. https://doi.org/10.1016/j.renene. 2014.01.002

23. Fleming P, Gebraad P, Lee S, et al. Evaluating techniques for redirecting turbine wakes using SOWFA. Renew Energy. 2014;70:211-218.

24. Xie S, Archer CL, Ghaisas N, Meneveau C. Benefits of collocating vertical-axis and horizontal-axis wind turbines in large wind farms. Wind Energy. 2017;20(1):45-62. https://doi.org/10.1002/we.1990

25. Calaf M, Meneveau C, Meyers J. Large eddy simulations of fully developed wind-turbine array boundary layers. Phys Fluids. 2010;22:15110. 26. Wu YT, Porté-Agel F. Large-eddy simulation of wind-turbine wakes: evaluation of turbine parametrisations. Bound-lay Meteorol. 2011;138:345-366. 27. Lu H, Porté-Agel F. Large-eddy simulation of a very large wind farm in a stable atmospheric boundary layer. Phys Fluids. 2011;23:65101.

28. Calaf M, Higgins C, Parlange MB. Large wind farms and the scalar flux over an heterogeneously rough land surface. Bound-Lay Meteorol. 2014;153(3):471-495.

29. Bou-Zeid E, Meneveau C, Parlange MB. A scale-dependent lagrangian dynamic model for large eddy simulation of complex turbulent flows. Phys

Fluids. 2005;17:25105.

30. Moeng CH. Large-eddy simulation model for the study of planetary boundary-layer turbulence. J Atmos Sci. 1984;41:2052-2062.

31. Albertson JD, Parlange MB. Natural integration of scalar fluxes from complex terrain. Advances in Water Resources. 1999;23(3):239-252. https://doi. org/10.1016/S0309-1708(99)00011-1

32. Stevens RJAM, Graham J, Meneveau C. A concurrent precursor inflow method for large eddy simulations and applications to finite length wind farms.

Renew Energy. 2014;68:46-50. http://dx.doi.org/10.1016/j.renene.2014.01.024

33. Stevens RJAM, Martínez-Tossas LA, Meneveau C. Comparison of wind farm large eddy simulations using actuator disk and actuator line models with wind tunnel experiments. Renew Energy. 2018;116:470-478. https://doi.org/10.1016/j.renene.2017.08.072

34. Munters W, Meneveau C, Meyers J. Shifted periodic boundary conditions for simulations of wall-bounded turbulent flows. Phys of Fluids. 2016;28:25112.

35. Chamorro LP, Porté-Agel F. Effects of thermal stability and incoming boundary-layer flow characteristics on wind-turbine wakes: a wind-tunnel study.

Bound-Lay Meteorol. 2010;136:515-533.

36. Chamorro LP, Porté-Agel F.. Turbulent flow inside and above a wind farm: a wind-tunnel study. Energies. 2011;4(11):1916-1936. https//:doi.org/10. 3390/en4111916

37. Lignarolo LEM, Mehta D, Stevens RJAM, et al. Validation of four LES and a vortex model against stereo-PIV measurements in the near wake of an actuator disc and a wind turbine. Renew Energy. 2016;94:510-523. http://dx.doi.org/10.1016/j.renene.2016.03.070

(14)

39. Wieringa J, Davenport AG, Grimmond CSB, Oke TR. New revision of davenport roughness classification. In: Proceedings of the 3rdEuropean & African

Conference on Wind Engineering; 2001; Eindhoven, The Netherlands.

40. Wu F. Simulation of turbulent flow inside and above wind farms: model validation and layout effects. Bound-Lay Meteorol. 2013;146(2):181-205. https://doi.org/10.1007/s10546-012-9757-y

41. Jiménez A, Crespo A, Migoya E, Garcia J. Advances in large-eddy simulation of a wind turbine wake. J Phys Conf Ser. 2007;75:12041.

42. Calaf M, Parlange MB, Meneveau C. Large eddy simulation study of scalar transport in fully developed wind-turbine array boundary layers. Phys Fluids. 2011;23(12):126603. https://doi.org/10.1063/1.3663376

43. Troldborg N, Sørensen JN, Mikkelsen R. Numerical simulations of wake characteristics of a wind turbine in uniform inflow. Wind Energy. 2010;13:86-99. 44. Troldborg N, Larsen GC, Madsen HA, Hansen KS, Sørensen JN, Mikkelsen R. Numerical simulations of wake interaction between two wind turbines

at various inflow conditions. Wind Energy. 2011;14:859-876.

45. Zhang M, Stevens RJAM. Exploring a better turbine layout in vertically staggered wind farms. J Phys Conf Ser. 2018;1037:72041. http://iopscience. iop.org/article/10.1088/1742-6596/1037/7/072041

46. Johnson KE, Fingersh LJ, Balas MJ, Pao LY. Methods for increasing region 2 power capture on a variable-speed wind turbine. ASME J Sol Energy Eng. 2004;126:1092-1100.

47. Porté-Agel F, Wu YT, Chen CH. A numerical study of the effects of wind direction on turbine wakes and power losses in a large wind farm. Energies. 2013;6:5297-5313.

48. Cal RB, Lebrón J, Castillo L, Kang HS, Meneveau C. Experimental study of the horizontally averaged flow structure in a model wind-turbine array boundary layer. J Renew Sustainable Energy. 2010;2(1):13106. https://doi.org/10.1063/1.3289735

49. VerHulst C, Meneveau C. Large eddy simulation study of the kinetic energy entrainment by energetic turbulent flow structures in large wind farms.

Phys Fluids. 2014;26:25113.

50. Stevens RJAM, Gayme DF, Meneveau C. Large eddy simulation studies of the effects of alignment and wind farm length. J Renew Sustainable Energy. 2014;6:23105. http://dx.doi.org/10.1063/1.4869568

51. Stevens RJAM, Gayme DF, Meneveau C. Effects of turbine spacing on the power output of extended wind-farms. Wind Energy. 2016;19:359-370. http://dx.doi.org/10.1002/we.1835

52. Barthelmie RJ, Pryor SC, Frandsen ST, et al. Quantifying the impact of wind turbine wakes on power output at offshore wind farms. J Atmos Oceanic

Tech. 2010;27(8):1302-1317. https://doi.org/10.1175/2010JTECHA1398.1

53. Markfort CD, Zhang W, Porté-Agel F. Turbulent flow and scalar transport through and over aligned and staggered wind farms. J of Turbulence. 2012;13(33):1-36.

54. VerHulst C, Meneveau C. Altering kinetic energy entrainment in large eddy simulations of large wind farms using unconventional wind turbine actuator forcing. Energies. 2015;8(1):370-386.

55. Allaerts D, Meyers J. Boundary-layer development and gravity waves in conventionally neutral wind farms. J Fluid Mech. 2017;814:95-130. 56. Allaerts D, Meyers J. Gravity waves and wind-farm efficiency in neutral and stable conditions. Bound-Lay Meteorol. 2018;166(2):269-299.

57. Stevens RJAM. Dependence of optimal wind turbine spacing on wind farm length. Wind Energy. 2016;19(4):651-663. https://doi.org/10.1002/we. 1857

58. Stevens RJAM, Gayme DF, Meneveau C. Using the coupled wake boundary layer model to evaluate the effect of turbulence intensity on wind farm performance. J Phys Conf Ser. 2015;625:12004. http://dx.doi.org/10.1088/1742-6596/625/1/012004

59. Stevens RJAM, Gayme DF, Meneveau C. Generalized coupled wake boundary layer model: applications and comparisons with field and LES data for two real wind farms. Wind Energy. 2016;19(11):2023-2040. http://dx.doi.org/10.1002/we.1966

60. Lissaman PBS. Energy effectiveness of arbitrary arrays of wind turbines. J of Energy. 1979;3(6):323-328.

61. Frandsen S. On the wind speed reduction in the center of large clusters of wind turbines. J Wind Eng and Ind Aerodyn. 1992;39:251-265. 62. Duckworth A, Barthelmie RJ. Investigation and validation of wind turbine wake models. Wind Eng. 2008;32:459-475.

63. Voutsinas S, Rados K, Zervos A. On the analysis of wake effects in wind parks. Wind Eng. 1990;14:204-219.

How to cite this article: Zhang M, Arendshorst MG, Stevens RJAM. Large eddy simulations of the effect of vertical staggering in large wind farms. Wind Energy. 2019;22:189–204.https://doi.org/10.1002/we.2278

APPENDIX A: JENSEN MODEL

In the Jensen wake model calculations presented in section 3.3, we assume that the wind turbine wakes grow linearly with downstream distance such that the wind velocity in the wakes is given by7,60

u(x) u0(z) = 1 −ΔUhub(x) u0(z) = ( 1 −1 − √ 1 − CT (1 + kwx∕R)2 ) = u0 ( 1 − 2a (1 + 2kwx∕D)2 ) , (A1)

with kw is the wake expansion coefficient, u0(z) is the incoming free stream velocity, R is the rotor radius, D is the turbine diameter, CT = 4a(1 −a) = 3∕4is the thrust coefficient with flow induction factor a= 1∕4, and x is the downstream distance with respect to the turbine.

(15)

velocity. Assuming that the former is on the order of the friction velocity, the ratio defining the wake decay parameter becomes

kw= 𝜅

ln(zh∕z0),

(A2) which we use to set the wake expansion coefficient for each turbine.15As in the vertically staggered wind farm, the hub height of the turbines varies between 60 and 150 m, this results (using𝜅 = 0.4and z0 = 0.1m) in kwvalues in range of0.0547to0.0625. Alternatively, one could set

the wake expansion coefficient using the Frandsen61model

kw=

0.5 ln(zh∕z0),

(A3) which would give kwvalues in range of0.0682to0.0782.

Various methods have been proposed to consider the wake-wake interaction in the Jensen model.2-4,62For example, using the superposition of squared velocity deficits according

Ui= u0− [ ∑ k (u0− uki)2 ]1∕2 , (A4)

as suggested by Katic et al11or by additionally taking into account the velocity deficit at each turbine location63

Ui= u0− [ ∑ k (uk− uki)2 ]1∕2 . (A5)

For the results presented in section 3.3, we assumed the latter. The normalized power of each turbine is then calculated by comparing the averaged cubed velocity at the location of each turbine with the corresponding value obtained for a turbine in the free stream. In order to account for the effect of the vertical staggering, we model the incoming free stream velocity

u0(z) u∗ =1 𝜅log ( z z0 ) . (A6)

Figure A1 compares the Jensen model predictions obtained using Equations A1, A2, A5, and A6 and without the use of ghost turbines, with LES results. The figure shows that the Jensen model predictions for the aligned wind farm cases do not capture the gradual recovery of the power output as function of downstream direction due to the vertical kinetic energy flux observed for these LES cases. For the vertically staggered wind farms, the Jensen model predicts a similar zigzag pattern as function of downstream direction as observed in the LES result. However, for the wind farm with the small turbines (Ds = 100m and zh = 100m), the Jensen model overpredicts the production by the tall turbines further

downstream. Figure A2 shows the effect of several modeling choices on the Jensen model predictions for the fully developed regime is limited, ie, for all modeling choices tested here, the Jensen model predicts an appreciable benefit of vertical staggering for the wind farm with small turbines (Ds = 100m and zh = 100m) and small streamwise spacings, while the LES results do not show this benefit of vertical staggering in the

fully developed regime of the wind farm.

FIGURE A1 Comparison of the normalized power output PPref,1obtained from LES and the Jensen model predictions as function of downstream position. Pref,1is the power output of the first turbine row in the reference aligned wind farm. Left column: smaller wind turbines with Ds = 100m, zh = 100m, and Hd = 0m,40m (case a0and case a4); right column: larger wind turbines with Dl = 150m, zh = 120m, and Hd = 0m,30m (case A0and case A3) [Colour figure can be viewed at wileyonlinelibrary.com]

(16)

FIGURE A2 Jensen model predictions for the benefit of vertical staggering in wind farms with the smaller turbines (Ds = 100m and

zh = 100m) as function of streamwise turbine spacing sxfor different values of Hd, see Figure 1. The base case presented in section 3.3 assumes

the use of a logarithmic profile Equation A6 for the incoming free stream velocity, no ground modeling, while the wake expansion coefficients for the turbines are set by Equation A2, and the wake-wake interaction is accounted for using Equation A5. A, Effect of incoming free stream velocity profile. B, Effect of ground modeling. When ground modeling is activated, image/ghost turbines are included at z = −zh. C, Effect of wake expansion coefficient and D, the effect of the wake superposition model [Colour figure can be viewed at wileyonlinelibrary.com]

Referenties

GERELATEERDE DOCUMENTEN

La fouille de la maison de fief d' Azy illustre la modestie des feudataires unis aux families seigneuriales successives du bail1'4ge de Chassepierre et le déclin

• A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the

As it is also possible to construct an LDA classi fier on data from other subjects, we compare the performance of the new method to rLDA trained on cross-subject data.. The

Hoewel we niet meer kunnen vast- stellen of Bijns haar argumenten opzettelijk zo heeft gerangschikt, is het vanuit argumentatietheoretisch perspectief wel te verklaren dat een

In chapter 1, when setting out the basic structure and the basic theoretical knowledge about the world-systems theory, it was explained that Wallerstein himself

Awareness, MyEndometriosisTeam, Endometriosis Research Center ...48 Table 15: Behçet’s disease Top 3 categories according to number of nodes for each page-like network

In latere fases die in het teken stonden van scherpstellen en details, was er vooral ‘low key beïnvloeding’ (P. Nederlof, persoonlijke communicatie, 21 april en 2 mei 2017)

In this work the process chain simulation consists of three manufacturing steps for the coating, drying and calendering, due to existing process models in the literature and