You are on page 1of 133

WIND ARRAY PERFORMANCE EVALUATION MODEL FOR LARGE WIND

FARMS AND WIND FARM LAYOUT OPTIMIZATION

by

SIMENG LI

Submitted in partial fulfillment of the requirements

For the degree of Doctor of Philosophy

Department of Mechanical and Aerospace Engineering

CASE WESTERN RESERVE UNIVERSITY

August, 2014

i
CASE WESTERN RESERVE UNIVERSITY
SCHOOL OF GRADUATE STUDIES

We hereby approve the thesis/dissertation of

Simeng Li

candidate for the degree of Doctor of Philosophy in Mechanical Engineering

Committee Chair

J. IWAN D. ALEXANDER

Committee Member

JAIKRISHNAN R. KADAMBI

Committee Member

PAUL BARNHART

Committee Member

DAVID H. MATTHIESEN

Date of Defense

July 2, 2014

*We also certify that written approval has been obtained

for any proprietary material contained therein

ii
Contents
Chapter 1 Introduction ......................................................................................................................... 1
1.1 Background ............................................................................................................................. 1
1.2 Wake measurements and wake model .................................................................................. 5
1.3 Wind farm layout optimization .............................................................................................. 6
1.4 Dissertation Outline................................................................................................................ 7
Chapter 2 Literature Review ................................................................................................................. 8
2.1 Wake models .......................................................................................................................... 8
2.1.1 Analytical Wake models ................................................................................................. 8
2.1.2 Numerical Wake model ................................................................................................ 10
2.1.3 Multiple wake models .................................................................................................. 12
2.2 Wind farm performance evaluation models ........................................................................ 13
2.3 Wind farm layout optimization problem .............................................................................. 15
Chapter 3 Objectives .......................................................................................................................... 18
3.1 Wind array performance evaluation .................................................................................... 18
3.2 Wind farm layout optimization ............................................................................................ 19
Chapter 4 Wake models ..................................................................................................................... 21
4.1 Kinematic Wake Models ....................................................................................................... 23
4.1.1 Jensen wake model ...................................................................................................... 24
4.1.2 Larsen model ................................................................................................................ 25
4.1.3 Frandsen analytical model............................................................................................ 26
4.2 Field Wake Models ............................................................................................................... 27
4.2.1 Ainslie wake model ....................................................................................................... 28
4.2.2 Three-dimensional field model .................................................................................... 29
4.3 Wake added turbulence models .......................................................................................... 31
Chapter 5 Methodology...................................................................................................................... 34
5.1 Large wind array performance evaluation model (LWAP) ................................................... 34
5.1.1 Multiple wake model .................................................................................................... 36
5.1.2 Effect of the wind array on the atmospheric boundary layer ...................................... 42
5.2 Wind array layout optimization model (WALOM)................................................................ 46
5.2.1 Wind array configuration set up................................................................................... 46
5.2.2 Wind data and distribution........................................................................................... 49

iii
5.2.3 Cost function ................................................................................................................ 55
5.2.4 Turbine power .............................................................................................................. 56
5.2.5 Wake effect evaluation................................................................................................. 57
5.2.6 Genetic Algorithm (GA) ................................................................................................ 58
Chapter 6 Results: wind farm layout evaluation model ..................................................................... 63
6.1 Wind speed evaluations at Horns Rev when wind direction is along turbine rows ............. 63
6.2 Turbine power evaluations at Horns Rev: wind direction parallel to turbine rows ............. 69
6.3 Power output predictions for turbines in the row at the Horns Rev and Nysted for a
representative wind speed and variable wind directions ................................................................ 73
Chapter 7 Results: wind farm layout optimization ............................................................................. 79
7.1 Extension of Mossetti’s approach ........................................................................................ 79
7.1.1 Case 1: Constant unidirectional wind ........................................................................... 80
7.1.2 Case 2: Constant wind speed with an equal probability variable wind direction ........ 85
7.1.3 Case 3: Variable Wind Speed with Variable Wind Direction ........................................ 87
7.1.4 Different Spacing Limits for Unidirectional Wind ......................................................... 89
7.1.5 Different Area Sizes for Unidirectional Wind ............................................................... 92
7.1.6 Constant Wind Speed with Variable Wind Direction and a circular site area .............. 95
7.1.7 Comparison of optimized layouts using Jensen wake model and Ainslie wake model 97
7.2 Horns Rev wind farm layout optimization.......................................................................... 100
7.2.1 Weibull data for 12 wind direction sectors ................................................................ 103
7.2.2 Wind data for 16 wind direction sectors .................................................................... 105
7.2.3 Wind data for 72 wind direction sectors .................................................................... 107
Chapter 8 Conclusions and future work ........................................................................................... 110
Bibliography ........................................................................................................................................ 113

iv
Tables
TABLE 5.1 WIND DATA EXAMPLE 51
TABLE 5.2 WEIBULL FACTORS FOR DIFFERENT WIND DIRECTIONS [77]. 54
TABLE 6.1 MAPE OF THE COMPUTED NORMALIZED WIND VELOCITY USING OBSERVED DATA FOR TWO WIND
DIRECTIONS AND TWO WIND SPEEDS REPORTED IN [60] AT HORNS REV. 69
TABLE 6.2 MAPE OF THE COMPUTED NORMALIZED TURBINE POWER USING COMPUTED PREDICTIONS AND
ACTUAL OBSERVATIONS AT HORNS REV [58]: WIND DIRECTION PARALLEL TO TURBINE ROWS. 72
TABLE 6.3 RMSD OF THE COMPUTED NORMALIZED POWER FOR VARIOUS WIND DIRECTIONS AND A WIND
SPEED OF 8 M/S AT HORNS REV WIND FARM [60] 78
TABLE 6.4 RMSD OF COMPUTED NORMALIZED POWER FOR VARIOUS WIND DIRECTIONS AND A WIND SPEED
OF 8 M/S AT NYSTED WIND FARM [60] 78
TABLE 7.1 RESULTS FROM PREVIOUS STUDY AND CURRENT STUDY: REPORTED AND RECOMPUTED 83
TABLE 7.2 RESULTS FROM PREVIOUS STUDY AND CURRENT STUDY: REPORTED AND RECOMPUTED 86
TABLE 7.3 RESULTS FROM PREVIOUS STUDY AND CURRENT STUDY: REPORTED AND RECOMPUTED 89
TABLE 7.4 TURBINES DISTRIBUTIONS IN OPTIMIZED LAYOUTS FOR ISOTROPIC WIND AND ROUND AREA 96
TABLE 7.4 VESTAS V80 THRUST COEFFICIENT AND POWER AS A FUNCTION OF WINS SPEED 101
TABLE 7.5 WIND DISTRIBUTION FOR HORNS REV 102
TABLE 7.6 TURBULENCE INTENSITIES FOR VARIABLE WIND SPEEDS 102
TABLE 7.7 LAYOUT PERFORMCANCE OF OPTIMIZATION RESULTS 109

v
Figures
FIGURE 1.1 AERIAL VIEW FROM THE SOUTHEAST OF WAKE CLOUDS AT HORNS REV ON FEBRUARY 12, 2008
[11]. 2
FIGURE 4.1 WAKE PROFILE DOWNSTREAM A TURBINE 21
FIGURE 4.2 TURBINES TAKEN INTO CONSIDERATION WHEN CALCULATING ADDED TURBULENCE 32
FIGURE 5.1 OVERLAPPED WAKES WHERE 𝐬𝟏=7D 38
FIGURE 5.2 WAKE DECAY CONSTANT, 𝒌′, AS A FUNCTION OF UPSTREAM TURBINE WIND SPEED DEFICIT 39
FIGURE 5.3 COMBINATION COEFFICIENT, C, AS A FUNCTION OF NORMALIZED DOWNSTREAM DISTANCE 40
FIGURE 5.4 NORMALIZED WIND VELOCITY CALCULATED USING MULTIPLE WAKE MODEL AND THE JENSEN
WAKE SINGLE MODEL. 41
FIGURE 5.5 PREDICTED FREE STREAM WIND SPEEDS AT TURBINE HEIGHT FOR THE HORNS REV WIND FARM
LAYOUT 45
FIGURE 5.6 WIND FARM BOUNDARY SET UP 47
FIGURE 5.7 HORNS REV ARRAY TURBINE LAYOUT COORDINATES 48
FIGURE 5.8 HORNS REV TURBINES RANKING FOR NORTH WIND 48
FIGURE 5.9 HORNS REV TURBINES RANKING FOR WEST WIND 49
FIGURE 5.10 MEAN WIND SPEED FREQUENCY DISTRIBUTION (WEIBULL SHAPE FACTOR, K, AND SCALE 52
FIGURE 5.11 WIND DIRECTION ROSE 53
FIGURE 5.12 COST OF WIND FARM VS. NUMBER OF TURBINES 55
FIGURE 5.13 VESTAS V80 2MW TURBINE POWER OUTPUT AND THRUST COEFFICIENT VS. WIND SPEED [78] 57
FIGURE 5.14 GENETIC ALGORITHM PROCESS 61
FIGURE 5.15 FLOWCHART OF GENETIC ALGORITHM OPTIMIZATION PROCESS 62
FIGURE 6.1 HORNS REV LAYOUT: CASE 1 OF 270° AND 7D SPACING, CASE 2 OF 222° AND 9.4 D SPACING 64
FIGURE 6.2 HORNS REV EVALUATION WIND DIRECTION 270° AND WIND SPEED 8.5 M/S +/- 0.5 M/S 65
FIGURE 6.3 HORNS REV EVALUATION WIND DIRECTION 270° AND WIND SPEED 12 M/S +/- 0.5 M/S 66
FIGURE 6.4 HORNS REV EVALUATION WIND DIRECTION 222° AND WIND SPEED 8.5 M/S +/- 0.5 M/S 67
FIGURE 6.5 HORNS REV EVALUATION WIND DIRECTION 222° AND WIND SPEED 12 M/S +/- 0.5 M/S 68
FIGURE 6.6 POWER CURVE FOR THE TURBINE AT HORNS REV 70
FIGURE 6.7 TURBINES POWER AT CASE 1 FOR WIND SPEED AT 8M/S AND DIRECTION 270° AT HORNS REV 70
FIGURE 6.8 TURBINES POWER AT CASE 1 FOR WIND SPEED AT 10M/S AND DIRECTION 270° AT HORNS REV 71
FIGURE 6.9 TURBINES POWER AT CASE 2 FOR WIND SPEED AT 8M/S AND DIRECTION 222° AT HORNS REV 71
FIGURE 6.10 HORNS REV ARRAY. EXACT ROW (ER=270°) OF TURBINES [60] 74
FIGURE 6.11 NYSTED ARRAY. EXACT ROW (ER=278°) OF TURBINES [60] 74
FIGURE 6.12 NORMALIZED POWER AT HORNS REV FOR THE FREE STREAM WIND SPEED OF 8 ± 0.5 M/S:
COMPARISON OF MODELS WITH OBSERVATIONS 75
FIGURE 6.13 NORMALIZED POWER AT NYSTED FOR THE FREE STREAM WIND SPEED OF 8 ± 0.5 M/S:
COMPARISON OF MODELS WITH OBSERVATIONS 76
FIGURE 7.1 WIND FARM AREA 82
FIGURE 7.2 WIND DISTRIBUTION FOR CASE 3 82
FIGURE 7.3 FITNESS VALUE OF DIFFERENT NUMBER OF TURBINES FOR CASE 1. 83
FIGURE 7.4 TURBINES PLACEMENT OF FOUR STUDIES FOR CASE 1: (A) MOSSETTI ET AL. [25] (B) GRADY ET AL.
[27] (C) MARMIDIS ET AL. [28] (D) MITTAL ET AL. [29] (E) WALOM 84
FIGURE 7.5 FITNESS VALUE OF DIFFERENT NUMBER OF TURBINES FOR CASE 2 85
FIGURE 7.6 TURBINES PLACEMENT FOR CASE 2: (A) MOSSETTI ET AL. [25] (B) GRADY ET AL. [27] (C) MITTAL ET
AL. [29] (D) WALOM 86

vi
FIGURE 7.7 FITNESS VALUE OF DIFFERENT NUMBER OF TURBINES FOR CASE 3 87
FIGURE 7.8 TURBINES PLACEMENT FOR CASE 3: (A) MOSSETTI ET AL. [25] (B) GRADY ET AL. [27] (C) MITTAL ET
AL. [29] (D) WALOM 88
FIGURE 7.9 FITNESS VALUES OF DIFFERENT SPACING LIMITS 90
FIGURE 7.10 EFFICIENCIES OF DIFFERENT SPACING LIMITS 90
FIGURE 7.11 OPTIMAL PLACEMENTS OF FOUR SPACING LIMITS FOR 40 TURBINES 91
FIGURE 7.12 FITNESS VALUES OF DIFFERENT AREA SIZES 93
FIGURE 7.13 EFFICIENCIES OF DIFFERENT AREA SIZES 93
FIGURE 7.14 OPTIMAL PLACEMENT OF FOUR AREA SIZES FOR 40 TURBINES 94
FIGURE 7.15 FITNESS VALUE AS A FUNCTION OF TURBINE NUMBER 95
FIGURE 7.17 OPTIMAL TURBINES LAYOUTS FOR A CIRCULAR SITE AREA 96
FIGURE 7.18 OPTIMAL TURBINES LAYOUTS FOR 48 TURBINES WITH CASE 1 UNIFORM ONE DIRECTION (FROM
NORTH TO SOUTH) WIND USING JENSEN WAKE MODEL AND AINSLIE WAKE MODEL. 98
FIGURE 7.19 OPTIMAL TURBINES LAYOUTS FOR 48 TURBINES WITH CASE 2 FOR VARIABLE WIND DIRECTION
AND A CONSTANT WIND SPEED USING JENSEN WAKE MODEL AND AINSLIE WAKE MODEL. 99
FIGURE 7.20 OPTIMAL TURBINES LAYOUTS FOR 48 TURBINES WITH CASE 3 FOR VARIABLE WIND DIRECTION
USING THE JENSEN WAKE AINSLIE WAKE MODELS. 99
FIGURE 7.21 WIND DIRECTION DISTRIBUTION FOR 12 DIRECTION SECTORS 103
FIGURE 7.22 OPTIMIZED LAYOUT FOR CASE 1 104
FIGURE 7.23 ANNUAL WIND POWER ROSE WITH 12 DIRECTION SECTORS 104
FIGURE 7.24 WIND DIRECTION DISCRETIZATION FOR 16 DIRECTION SECTORS 106
FIGURE 7.25 ANNUAL POWER COMPARISON OF 16 DIRECTION SECTORS 106
FIGURE 7.26 WIND DIRECTION DISCRETIZATION FOR 72 DIRECTION SECTORS 107
FIGURE 7.27 OPTIMIZED LAYOUT FOR 72 DIRECTION SECTORS 108

vii
Nomenclature
𝐴 Rotor disc area
𝐴𝑛 The nth wake area
𝐴𝑟𝑛 Rotor area of the nth turbine
𝐴𝑐 The Weibull scale factor
𝐴𝑗 The Weibull scale factor of the jth wind direction
𝑎 Turbine induction factor
𝑏 Wake lateral width
𝐶 Combination factor
𝐶𝑇 Wind turbine thrust coefficient
𝐶𝑡 The distributed thrust coefficient
𝑐1 Constants in Larsen wake model
𝑐𝑚𝑤 The relative mean wind speed in the wake
𝑐𝑤𝑓 The flow speed deficit in the infinitely large wind farm
cost Total cost of the wind farm
𝐷 Turbine rotor diameter
𝐷𝑒𝑓𝑓 The effective rotor diameter
𝐷𝑀 Centerline velocity deficit
𝐷𝑚𝑖 The initial centerline velocity deficit
𝐷𝑟 The expanded downstream rotor diameter
𝐷𝑠 Empirical distance constant
𝐷𝑤 Wake width
ℎ Height of the internal boundary layer
ℎ𝐻 Turbine hub height
𝐼𝑎 The ambient turbulence intensity
𝐼𝑎′ The ambient turbulence intensity in a turbine wake
𝐼𝑇 Maximum center wake turbulence intensity
𝐼𝑤 Turbulence intensity in the wake
𝐾 The Weibull shape factor
𝐾𝑗 The Weibull shape factor of the jth wind direction
𝑘 Wake decay constant
𝑘′ Wake decay coefficient
𝑁 Number of turbines in the wind farm
𝑃 Power output of a turbine
𝑃𝑎𝑣𝑎𝑖𝑙 Power available from the wind
𝑃𝑜𝑏𝑠𝑒𝑟𝑣𝑒𝑑 Observed turbine power
𝑃𝑝𝑟𝑒𝑑𝑖𝑒𝑐𝑡𝑒𝑑 Predicted turbine power
𝑃𝑡𝑜𝑡𝑎𝑙 Total wind farm power

viii
𝑅9.5 The wake radius at 9.5 rotor diameters downstream the turbine
𝑅𝑤 Rotor wake radius
𝑠 Normalized downstream distance by turbine rotor diameter
𝑠𝑖 Normalized downstream distance from the ith turbine
𝑠𝑐 Crosswind turbine spacing
𝑠𝑑 Downwind turbine spacing
𝑠𝑓 Turbine spacing
𝑈′ The relative wake velocity deficit
𝑈1 Free stream velocity at turbine height in the internal boundary layer
𝑈𝑐 Normalized centerline velocity deficit
𝑈𝑖𝑛𝑖𝑡𝑖𝑎𝑙 The initial wake velocity deficit
𝑈0 Free stream wind velocity
𝑈0𝑗 Free stream wind velocity of the jth wind direction

𝑈 The mean wind speed
∆𝑈 Velocity deficit in the wake
𝑢 Wind velocity in a turbine wake
𝑢𝑖 Wind velocity calculated by the wake model in the ith turbine wake
𝑢𝑖′ Wind velocity calculated by the single wake model in the ith turbine wake
𝑢𝑜𝑏𝑠𝑒𝑟𝑣𝑒𝑑 Observed wind speed
𝑢𝑝𝑟𝑒𝑑𝑖𝑐𝑡𝑒𝑑 Predicted wind speed
𝑢∗0 Friction velocity in the atmospheric boundary layer
𝑢∗1 Friction velocity in the internal boundary layer
𝑥0 The position of the rotor respected to the applied coordinate system
𝑧0 Offshore roughness
𝑧00 Wind farm roughness
𝛼 Constant related to the trust coefficient
𝛽 Wake expansion parameter
ε𝑣 Eddy viscosity
𝜃𝑗 The jth wind direction
𝜅 Von Karman Constant
𝜌 Air density
𝜎𝑈 The standard deviation of the wind speed in the wake
𝜎𝑦 Standard deviation of wind velocity in y direction
𝜎𝑧 Standard deviation of wind velocity in z direction

ix
Abbreviations
CFD Computational Fluid Dynamics
ENDOW EfficieNt Development of Offshore WindFarms
ECN Energy Research Centre of the Netherlands
ER Exact Row
EWTS European Wind Turbine Standards
FLaP Farm Layout Program
GA Genetic Algorithms
GH Garrad Hassan
IBL Internal Boundary Layer
KAMM Karlsruhe Atmospheric Mesoscale Model
LWAP Large Wind Array Performance Evaluation Model
MAPE Mean Absolute Percentage Error
MC Monte Carlo
NTUA National Technical University of Athens
RGU Robert Gordon University
RMSD Root Mean Square Deviation
SAR Synthetic Aperture Radar
SODAR Sonic Detection and Ranging
WALOM Wind Array Layout Optimization Model
WAsP Wind Atlas Analysis and Application Program
WFOG Wind Farm Optimization using a Genetic Algorithm

x
Acknowledgements

Many people have contributed in one way or another to the completion of this work
and, while I cannot name them all, I wish to express my deep gratitude to each of them.
My first gratitude must go to my advisor, Dr. J. Iwan D. Alexander. He patiently
provided the vision, encouragement and advice necessary for me to proceed through the
doctoral program and complete my dissertation. I want to thank Iwan for his unflagging
encouragement and serving as a role model to me as a junior member of academia. He has
been a strong and supportive adviser to me throughout my graduate school career. In
addition, he has always given me a great freedom to work independently.
Special thanks to my committee, Professors Jaikrishnan R. Kadambi, Paul Barnhart
and David H. Matthiesen for their support, guidance, helpful suggestions and patience
throughout the course of my research. Their guidance has served me well and I owe them
my great appreciation.
I am heartily thankful to Hui Yi, for her help, support and encouragement
throughout my pursuit for the doctoral degree. I feel fortunate to have met you and cherish
the time we spent together in US.
Also, thanks to Professor Joseph Prahl for his constant help and encouragement on
my study and research. Thank you, Professor Bo Li, for your advice on my dissertation and
journal papers. I would also like to thank my colleagues, Yng-Ru Chen, Ying Chen, Xinyou Ke
and department assistant Sheila Campbell. I enjoy the time studying and working with you.
Finally, I would also like to express my gratitude to my parents for their care and
support. Also thanks to my friends Jinxia Guo, Hua Zhou, Lin Chen and Zhe Yang for their
support and encouragement.
This work was partially supported by the Department of Energy (Award Numbers
EE0000275 and DE-EE0005379). I would also like to acknowledge support from Case
Western Reserve University’s Great Lake Energy Institute.

xi
WIND ARRAY PERFORMANCE EVALUATION MODEL FOR LARGE WIND
FARMS AND WIND FARM LAYOUT OPTIMIZATION

Abstract

by

SIMENG LI

The grouping of wind turbines in arrays introduces two major issues: (1) reduced power

production caused by wake wind speed deficits and (2) increased dynamic loads on the

blades caused by higher turbulence levels. Depending on the layout and local wind

conditions, the drop in power production of downstream turbines can easily reach 40% of

the upstream turbines in fully developed wake conditions. These power drops across arrays

arise due to wake wind speed deficits. Even when averaged over different wind directions,

drops in power production of 8% (onshore arrays), and 12% (offshore arrays) have been

recorded. In this dissertation, a large wind array performance evaluation model (LWAP) to

evaluate wake effects in large wind farms is developed. The model accounts for multiple

wake interactions and the effect on the vertical wind profile in the atmosphere boundary

layer by the wind farm itself. The model predicts wind speed deficits at each turbine and for

specific turbine power curves and assesses power for individual turbines and for the entire

wind farm. The calculation method converges within seconds for a large wind farm

evaluation. To assess the efficacy of the wake model, measured wind speed deficits and

turbine power deficits along two wind directions and wind turbine rows in the Horns Rev

xii
wind farm were compared with deficits calculated using the model. The mean absolute

percentage error is around 2% on average in wind speed evaluation and around 4% on

average in wind turbine power evaluation. Case studies predicting row-wise power deficits

of turbines arrays in Horns Rev and Nysted wind farms on multiple wind directions were

compared to observations. LWAP exhibits the same accuracy on power deficit evaluation as

with the CFD based models such as WindFarmer, WakeFarm and NTUA and performs better

than the WAsP Park model. The computing time to process an entire full wind farm (e.g.,

Horns Rev) is on the order of a few seconds, significantly less than the CFD based models. In

addition, a wind array layout optimization model (WALOM) is proposed to simulate,

evaluate and optimize wind array performance for real wind farm site. Results of optimized

wind array layouts are obtained and analyzed on case studies of multiple wind distributions

conditions and site conditions. It is found that the optimized results are affected by factors

such as wind distribution, wind data resolution, wake model and wind farm site conditions.

xiii
Chapter 1 Introduction

1.1 Background

Wind energy is the fastest growing source of electricity and one of the fastest

growing markets in the world. At the end of 2011, worldwide wind power capacity reached

238 GW, doubled in three years [1, 2]. Wind power is renewable, clean, worldwide and

using little land [3]. In the year of 2010, it was reported that 430 TWh was generated by

wind power, which is about 2.5% of worldwide electricity usage [4, 5]. It is expected to

reach 3.35% by 2013 and 8% by 2018 [6, 7]. For the U.S. to reach 54 GW of offshore wind

energy by 2030 [8], 1000’s of wind turbines (typical offshore array sizes today range from 40

-100 turbines) will need to be installed along the Atlantic, Pacific, and Great Lakes coasts.

The cost of these developments must be competitive with traditional on-shore generation,

requiring new and optimized designs now.

Wind turbines, which transform wind power into electricity, are usually grouped into

large wind farms, also called wind arrays, for reason of economies, such as lower

transportation, installation, maintenance and land costs. However, grouping turbines causes

a reduction of the power produced due to wake effects. When wind turbine extracts energy

from the wind, it produces a wake of wind velocity deficit downwind the turbine, so that the

power extracted by downstream turbines is reduced. Nowadays, a large wind farm may

consist of several hundred wind turbines, so that wake effects are unavoidable. Recent

studies found that the average power losses due to wake in a wind array are in the order of

1
10-20% [8]. In addition, the effect of increased fatigue loads of wind turbines operating in

wakes leads to a shorter turbine lifetime [9, 10].

Figure 1.1 Aerial view from the Southeast of wake clouds at Horns Rev on February 12, 2008 [11].

Figure 1 shows a famous photograph of the Horns Rev wind farm in Denmark. It

illustrates the wake by wind turbines and wake interactions in the large wind farm. During a

2
previous study [12] for this wind case, power losses of wind turbines downstream in the

wind farm reach 40%.

The ability to accurately quantify power losses associated with wind turbine wakes is

an important aspect of wind farm design analysis. One source of uncertainty in estimating

these losses is the effects of interacting wakes within the wind farm. Flow in the wake from

a wind turbine is characterized by momentum or velocity deficits and increased turbulence

levels that can adversely affect the performance of wind turbines situated downstream.

Accurate prediction of wind velocity deficit and wind turbine power deficit downstream of

wind turbines is crucial to the evaluation of wind farm layout. Specifically this reduction in

performance is manifested by subpar power output due to decreased wind speeds as well

as decreased efficiency due to fluctuating loads on the blades and tower.

It is also important that a detailed understanding is developed of the relevant

aspects of turbine placement within a wind farm and how that placement affects the wakes

and, thus, the expected power losses from a given wind distribution. Reducing wake losses,

or even reduce uncertainties in predicting power losses from wakes, contributes to the

overall goal of reducing the cost of wind-farm operation while maximizing power produced.

Accurate prediction of wind velocity deficit downstream of the wind turbines is crucial to

the estimation of the power output and loading of a wind farm.

The spatial configuration of wind turbines in a large scale wind power plant will

affect the overall performance of the plant and, thus, be an important contributing factor to

cost-effectiveness over the long-term. The grouping of turbines in arrays introduces two

major issues: (1) reduced power production caused by wake velocity deficits and (2)

3
increased dynamic loads on the blades caused by higher turbulence levels. Depending on

the layout and local wind conditions, the drop in power production of downstream turbines

can easily reach 40% of the upstream turbines in fully developed wake conditions. These

power drops across arrays arise due to wake velocity deficits. Even when averaged over

different wind directions, drops in power production of 8% (onshore arrays), and 12%

(offshore arrays) have been recorded (Barthelmie et al. [13]).

The land requirement for wind turbines is roughly 10 hectares (2.78 acres) per MW.

To date, with a few exceptions, most of the rationale behind spacing is intuitive, e.g., Patel

[14] proposed 8-12 rotor diameters windward 1.5-3 diameters crosswind. This was

suggested to be inefficient by Ammara [15] who proposed a denser staggered siting scheme

that would produce the same power but use less land. The placement of wind turbines in

large wind turbine arrays or wind fields will, thus, affect the overall power generation

characteristics of the wind field. If turbines in large arrays are not sited to attempt to

maximize performance while keeping within other practical site-particular constraints that

may be imposed (for example, with respect to wake interference between turbines and

with respect to the prevailing winds) then sub-par performance of the individual turbines

within the field may result. A good example of this is the Horns Rev (160 MW) wind field off

the coast of Denmark. Certainly, one of the most heavily studied wind field Horns Rev has

exposed limitations of current methods that are employed to determine turbine placement

in large wind turbine arrays [16-18].

4
1.2 Wake measurements and wake model

The measurement and characterization of wind turbine wakes, while relatively

straight forward under certain controlled laboratory conditions, i.e., scaled wind turbine

models in wind tunnels or small turbines in large wind tunnels is more complicated in the

field as wakes are spatio-temporally variable phenomena shifting with the direction of the

wind and not amenable to meteorological measurements on a long term basis without

excessive instrumentation [19]. Erection and maintenance of meteorological towers

offshore is costly and so wake measurements are typically limited to measurements

downstream of the prevailing wind direction. The Danish wind turbine group at RISO has

made the most advances in wake measurements. Three meteorological towers erected in

the vicinity of eleven turbines at the Vindeby offshore farm have permitted simultaneous

measurements of wind speed in the free stream and wake for several wind directions [20].

Measurements of velocity profiles at different distances downstream from the turbines

were also obtained by Barthelmie et al. [21] using ship-borne Sonic Detection and Ranging

(SODAR). Types of wake measurements were also discussed by Barthelmie et al. [8]. Wake

measurements have also been obtained from satellite SAR measurements by Christiansen

and Hasager [22].

Wind turbine wakes have been studied for two decades and various models have

been developed. These models can be divided into two main categories, namely, analytical

wake models and computational or numerical wake models. An analytical wake model

characterizes the velocity in a wake by a set of analytical expressions whereas in

computational wake models, fluid flow equations, whether simplified or not, must be solved

5
to obtain the wake velocity field. Analytical models have advantages from the viewpoint of

evaluating and designing wind farm performance because of its simplicity and

computational speed [17].

As wakes develop downstream, they interact with the atmospheric boundary layer

as well as with other wakes and are also affected by variable surface terrain. The extent of

the wake and the way in which the wake evolves is dependent on a number of factors.

These include the wind conditions (speed and turbulence intensity) entering the wind field

(the so-called inlet conditions), the terrain, the interaction of the wake with the

atmospheric boundary layer [23], the turbulence developed in the wake (wake-added

turbulence) and the characteristics of the turbine (size, blade design, hub height, etc.).

1.3 Wind farm layout optimization

At present, in existing wind arrays, turbines are often organized in identical rows

that are separated by a convenient spacing, normally of 6-10 rotor diameters [24]. As the

spacing between turbines increases, the wake losses decrease, but results in higher inter-

turbine cabling and land costs. The problem then is to balance these competing effects by

determining the distance from shore and the turbine spacing that results in lowest possible

wake effects.

Several attempts have been made to optimize turbine placement in wind arrays and

showed that irregular arrays result in a higher energy layout than regular grids [25-37]. One

of the earliest was by Mossetti et al. [25] who developed optimal placing schemes based on

Genetic Algorithms (GA). It seems to be a good way to go and could be optimized to

6
constrain placement to maximize power output, minimize surface area occupied by the

wind farm and minimize cost or maximize asset lifetimes as well as factoring in spatially

variable operating points for individual or groups of turbines. Mossetti [25] also applied

Jensen wake model [26] to evaluate the turbine wake. Many relevant researchers that used

the similar method to optimize this problem in the wind array can be found in the literature

[27-31]. However, the models in these studies for optimal wind turbine placement

proposed to date are strikingly simple and lack a robust wake model capable of accounting

for inter-turbine interactions. In addition, changes in operating points on the turbines

power curves according to turbulence intensity etc., are not accounted for.

1.4 Dissertation Outline

The dissertation is organized as follows: A review of previous work on wind turbine wake

modelling and wind farm layout optimization is presented in Chapter 2. A brief review of

optimization methods is also given. Chapter 3 defines and discusses the objectives of this

dissertation. Wind turbine wake and wind farm related models form the core part of the

work and the optimization modeling and are discussed in Chapter 4. Chapter 5 describes the

methodology of the wind array performance modeling. In Chapter 6, the wind array

performance evaluation model is verified against actual data obtained from Horns Rev wind

array and Nysted wind array. The application of the optimization model to two wind farm

layout case studies is described in Chapter 7. In addition, factors affecting the optimization

are analyzed. Finally, the conclusions and ideas for future work are presented in Chapter 8.

7
Chapter 2 Literature Review

This chapter provides the literature review and relevant background for this

dissertation. Section 2.1 reviews wake models that have been widely used in wind turbine

and wind farm researches. In section 2.2, models for wind farm performance evaluation and

wind farm design analysis are reviewed. Section 2.3 provides the background for the wind

farm layout optimization problem.

2.1 Wake models

Wake models can be divided into two main categories, namely, analytical wake

models and computational or numerical wake models. An analytical wake model

characterizes the velocity in a wake by a set of analytical expressions whereas in

computational wake models, fluid flow equations, whether simplified or not, must be solved

to obtain the wake velocity field.

2.1.1 Analytical Wake models

Analytical models are first introduced by Lanchester and Betz [38, 39], who derived

the principles of conservation of mass and momentum of the wind flow through an

idealized actuator disk that extracts energy from the wind.

Jensen wake model [26] treated the wake behind the wind turbine as a turbulent

wake which ignores the contribution of vortex shedding that is significant only in the near

wake region. The wake model is thus derived by conserving momentum downstream of

8
wind turbine. The velocity in the wake is given as a function of downstream distance and it

is assumed that the wake expands linearly downstream of the wind turbine. Jensen also

proposed that when two wakes interact, the resultant kinetic energy deficit is equal to the

sum of the kinetic energy deficits of the individual wakes at that point.

The work of Katic et al. [40] expanded the previous work of Jensen and assumes the

wake velocity profile is constant or a ‘top hat’ profile. This assumption is justified by the fact

that the purpose of the model is to estimate the energy content within the wind field as

seen by downwind turbines rather than accurately describing the spatial variation velocity

field. Notably, the Katic model is currently the basis of the wake model used in WAsP [41].

WAsP is a wind climate and turbine power prediction software developed by Risø and is

widely used.

Frandsen’s kinematic model based on work completed by Larsen et al. [42], and

currently included in the European Wind Turbine Standard II, is based on the ‘classical’ wake

theory outlined by Schlichting [43]. This model proposes a semi-analytical solution for

predicting the velocity deficit within a wake in which a set of simple empirical relations are

used to predict the turbulence intensity and the turbulence length scale. It assumes an

axisymmetric wake within which the velocity deficit decays with downstream distance to

the power of 2/3 and the turbulence intensity decays to the power of 1/3. It also suggests

that the wake width increases with the downstream distance to the power of 1/3.

Ishihara et al. [44] developed an analytical wake model by taking the effect of

turbulence on the rate of recovery into account. They used similarity theory for the velocity

profile and defined wake recovery as a function of ambient turbulence and turbine

9
generated turbulence. They calculated results for both offshore and onshore conditions and

also at both high loading and low loading of wind turbine. These results compared well with

experimental data obtained using a 1/100 scale model of Mitsubishi MWT-1000 wind

turbine in a wind tunnel. The scale model used surface roughness models upstream of the

wind turbine to simulate onshore conditions and a smooth upstream surface to simulate

offshore conditions.

Werle [45] proposed a three part wake model: an exact model for the inviscid near

wake region; Prandtl’s turbulent shear layer mixing solution for the intermediate wake and

a far wake model based on the classical Prandtl/Swain axisymmetric wake analysis. No

comparisons of the model with actual data have been published to date.

Lissaman’s kinematic wake model [46] predicts the effects of individual turbine

wakes, based on self-similar velocity deficit profiles. The work involved both experimental

and theoretical components related to co-flowing jets. The results give a full definition of

the velocity profile within a wake and facilitate the prediction of the velocity deficit for a

given wake radius. The growth of the wake depends on the ambient, free stream turbulence,

the shear generated turbulence and the turbulence created by the turbine. The maximum

velocity deficit at each downwind position is found via a control volume momentum

balance, with the initial velocity deficit calculated using the thrust coefficient of the turbine.

2.1.2 Numerical Wake model

There are a number of numerical wake models that range in complexity, and are

used by the wind turbine industry. Early work by Templin [47] and Newman [48] described

10
the effects of turbines as distributed roughness elements and were developed further by

Bossanyi et al. [49], Frandsen [50] and Emeis and Frandsen [51]. These models are useful

when predicting the effects of large wind farms on wind flow [52]. In another type of wake

model wind turbines are modeled as roughness elements [50, 52]. Frandsen combined the

drag from turbine with surface drag to get the total drag. The limitation of these types of

models is that the calculated total roughness is independent of wind direction and these

models are best suited for predicting overall effects of large wind farms on wind

characteristics.

Crespo et al. [52] carried out an extensive survey of different modeling methods for

wind turbine wakes. Apart from surveying various analytical wake models (discussed above)

she reported the computational wake model UPMWAKE to be one of the best after

comparing various models with wind tunnel measurements.

Field models (also known as implicit models), in contrast to kinematic models,

calculate the flow at every position throughout the wake field. The original work of this type

was completed by Sforza et al. [53] and describes a wake using the linearized conservation

of momentum equation in the direction of the free stream flow. The wake is assumed to

have constant advective velocity, constant eddy diffusivity and a wake shape described by a

parabolic approximation.

Other field models have been developed by Taylor [54], Liu et al. [55], Crespo et al.

[56], Ainslie [57] and, more recently, Magnusson [58]. These models involve solution of a

simplified version of the Reynolds averaged Navier–Stokes flow equations. They employed

11
an ‘eddy viscosity’ turbulence model to model the turbulent mixing contributions from the

shear layer and the ambient free stream turbulence.

2.1.3 Multiple wake models

One multiple wake calculation method was described by Mossetti et al. [25].

Multiple wakes were accounted for by simply assuming that the kinetic energy deficit of a

mixed wake was equal to the sum of energy deficits. The velocity downstream of n turbines

was then calculated using the following expression


𝑛
𝑢 2 𝑢𝑖
�1 − � = �(1 − )2 2.1
𝑈0 𝑈0
𝑖=1

where 𝑢 is the wind velocity for the turbine in the wake, 𝑈0 is the free stream wind velocity,

and 𝑢𝑖 is the wind velocity calculated in the ith turbine wake.

This approach to accounting for multiple wakes was also applied in other studies

[27-31]. However, Li et al. [32] showed that predictions with this method were not in

agreement to the measured wind speed deficits at Horns Rev array.

Another method was developed by Frandsen et al. [15] who presented a wake

expansion model developed by taking the control volume analysis of the momentum flow.

This model was used in WAsP engineering model for the wake evaluation [41]. The wind

speed at the n+1 downwind turbine can be found from

𝐴𝑛 𝐶𝑇 𝐴𝑛+1
𝑈𝑛+1 = 1 − �1 − 𝑈𝑛 �1 − � 𝐴𝑟𝑛 � 2.2
𝐴𝑛+1 2 𝐴𝑛

12
where 𝐶𝑇 is the thrust coefficient, 𝐴𝑛 is the nth wake area and 𝐴𝑟𝑛 is the rotor area of the

nth turbine.

The limitation of this model is that it applies only to a single row of equally spaced

turbines aligned parallel to the wind direction. To describe real wind farms an extension of

the approach to account for an irregular wind array layout is needed. In addition, any

evaluation of multiple wake effects on the performance of the wind farm needs to consider

a representative description of the distribution of expected wind speeds and directions for a

given geographic location, i.e., a large number of discrete wind directions and wind speeds.

Finally, the required computational time must to be sufficiently low to allow for efficient

computation of wind farm layout optimizations.

2.2 Wind farm performance evaluation models

In this section, four models for the wind farm performance evaluation model are

introduced. They are the Wind Atlas Analysis and Application Program (WAsP), the GH

WindFarmer, the ECN’s WAKEFARM and the NTUA model [41, 59-65].

WAsP is based on a linearized model used in the European Wind Atlas. The WAsP

program [41] is a series of models that developed from site measured wind data at a

generalized wind climate and hence is restricted to location specific descriptors of the flow

climate. In terms of wind farm modelling, the Park wake model [40] is used in the

commercial version. A new wake model with ‘Mosaic tile’ is being developed for use within

WAsP which is described by Rathmann et al. [59]. The program utilizes the observed wind

data by fitting it to a two parameter Weibulll distribution. The main advantage of the WAsP

13
program is that it is fast and robust. It is reported that for running a full wind farm

simulation as Horns Rev wind farm, WAsP takes the order of minutes [60]. However, in the

model, turbulence is not specifically treated for the wind farm.

Garrad Hassan (GH) WindFarmer is a CFD based model [61]. It applies an empirical

representation of the wind turbine wake developed by Ainslie [57]. Empirical expressions

are applied to model turbulence intensities in the turbine wake and the superposition of

multiple wakes. Multiple wakes are determined by a consecutive downstream modelling of

individual wakes. In addition, for very large wind farms, the change of the vertical wind

profile as a result of the presence of wind turbines is modeled [62]. The computation time

for a full wind farm simulation for this model was reported as minutes [60].

ECN’s WAKEFARM model is developed from the UPMWAKE code and was developed

by the Universidad Polytecnica de Madrid [63]. This model is similar to the GH WindFarmer

model, but is extended to 3D. The wake model is computed using a 3D parabolized Navier-

Stokes code and applies a k-epsilon turbulence model. The parabolisation of equations

leads to an efficient calculation procedure, but for the near wake, an empirical velocity

profile is required to for initialization. The axial pressure gradients are assumed as external

forces and enforce the flow to decelerate. Because it computes a 3D flow filed for the

simulation of wind farms such as Horns Rev, several hours are needed to converge [60].

The NTUA CFD model is developed by National Technical University of Athens and is

based on the 3D Reynolds averaged incompressible Navier-Stokes equations [64, 65]. The

model applies the k-epsilon turbulence closure model and accommodates wind turbines

embedded in the grid as momentum sinks representing the thrust force applied on the

14
rotor disk. This model needs much more computation than models discussed above. It is

reported that it requires a time scale of days for the Horns Rev wind farm simulation [60].

2.3 Wind farm layout optimization problem

The spatial configuration of wind turbines in a large scale wind power plant will

affect the overall performance of the plant and, thus, be an important contributing factor to

its cost-effectiveness over the long-term. The placement of wind turbines in large wind

turbine arrays or wind fields will affect the overall power generation characteristics of the

wind field. If turbines in large arrays are not sited to attempt to maximize performance

while keeping within other practical site-particular constraints that may be imposed (for

example, with respect to wake interference between turbines and with respect to the

prevailing winds) then sub-par performance of the individual turbines within the field, and,

thus, the field itself, may result.

The land requirement for wind turbines is roughly 10 hectares (2.78 acres) per MW.

To date, with a few exceptions, most of the rationale behind spacing is intuitive e.g., Patel

[17] proposed 8-12 rotor diameters windward 1.5-3 diameters crosswind. This was

suggested to be inefficient by Ammara [18] who proposed a denser staggered siting scheme

that would produce the same power but use less land.

Most approaches to placement of wind turbine within large arrays are based on

simplified wake models that account for different levels of interaction between turbines. In

recent work, particularly that by Frandsen, there have also been attempts to bring in more

meteorological based models [15] (in this case the Karlsruhe Atmospheric Mesoscale Model,

15
KAMM) to assist in assessing the efficacy of given wind turbine array configurations. When

compared to Horns Rev data, this was found to lead to improvements in predictive ability

but the reader is cautioned that that this approach may not translate well to different

locations and conditions.

Other researchers have attempted to optimize the placement of wind turbines in

wind farms. The first (published) attempt was made by Mossetti [25] who used genetic

algorithm for optimization. Mossetti’s idea was to develop optimal placing schemes based

on genetic algorithms. This seems to be a good way to go and could be optimized to

constrain placement to maximize power output, minimize surface area occupied by the

wind farm and minimize cost or maximize asset lifetimes as well as factoring in spatially

variable operating points for individual or groups of turbines. The key ‘physical’ ingredient

in Mossetti’s model was the wake model. He used utilized Jensen’s model [26]. Three

different cases were analyzed.

The same problem was tackled by Grady et al. [27] who improved upon Mossetti’s

analysis by accounting for non-uniform and variable winds. From their study it is apparent

that there is great sensitivity to the assumptions made in the analysis. Grady showed that

Mossetti’s results were not optimum and gave improved results. However, there are some

inconsistencies in the reported results.

Marmidis [28] attempted the problem of optimal placement using a Monte Carlo

(MC) method for optimization. He analyzed only one case out of three cases analyzed by

Mossetti and Grady. The biggest problem with his adaptation of the MC approach is the size

of the parameter space to be scanned and the slow approach to a minimum.

16
Wan et al. [34] improved previous works by using a Weibull function to describe the

probability of wind speed distribution and an improved turbine speed-power curve to

estimate turbine power output.

Kusiak et al. [30] extended this problem by using a multi-objective evolutionary

strategy algorithm. They achieved the optimization problem which maximized the expected

energy generation, and also minimizes the constraint violations.

Mittal et al. [29] developed a code Wind Farm Optimization using a Genetic

Algorithm (WFOG) for optimizing the placement of wind turbines in large wind farms by

using Jensen wake model [26] and the Fuga wake model developed by Ishihara et al. [44].

Mittal also used a refined grid for the possible turbine position. It allowed more flexibility in

the placement of wind turbines. It was reported that the Fuga wake model estimated the

velocity in the wake more accurately than the Jensen model.

Chen et al. [37] focused on optimizing the wind farm layout of turbines with

different hub height and under the uncertainty with landowner’s financial and noise

concerns. Their work improved transparency-of-information that can potentially affect the

negotiation process between developers and landowners during early wind farm planning

for the onshore site.

17
Chapter 3 Objectives

The overall objective of this research is to create a simulation tool suitable for

characterization of the in-situ operating environment within an offshore wind farm. The

tool will supply quantitative information necessary for wind turbine and wind farm design,

and power production estimates, and also optimize wind array layout to maximize power

production.

3.1 Wind array performance evaluation

Accurately predicting the wind and is a necessary part of assessing the power

production potential of a given layout of turbines in the wind array area and is crucial for

the success of any wind farm project. For multi-turbine arrays, this includes modeling of

wind distribution, power curve and wake losses caused by one or more upwind turbines.

Reducing wake losses, or even reduce uncertainties in predicting power losses from wakes,

contributes to the overall goal of reducing the cost of wind-array operation while

maximizing power produced.

In this dissertation the goal is to develop a wind array performance evaluation

model which is fast and suitable for the large wind farm design. A multiple wake model was

developed by only considering the wake of the nearest upstream turbine and its operating

conditions which are affected by other upstream turbine wakes. The model was developed

from the flow momentum theory of the Jensen wake model [26], see also Chapter 4. This

approach requires much less calculation than two multiple wake models discussed above

18
and can be easily applied to regular or irregular wind array layout for multiple wind

directions.

3.2 Wind farm layout optimization

The models for optimal wind turbine placement proposed to date are strikingly simple

and lack a robust wake model capable of accounting for the effects of wakes from neighboring

turbines. In addition, changes in operating points on the turbine power curve according to

turbulence intensity etc. are also not accounted for. The wind distribution forwarded by

Mossetti et al. [25] only considered three different wind speeds and assumed an

oversimplified angular distribution. The potential for improvement is high if only by applying

actual observed wind data with different real site and applying simultaneous consideration

of how factors such as wind distribution, wake model, optimization algorithms and site area

limitation will affect optimization results.

The wind farm layout optimization analyzes turbine positioning, power production,

turbine layout and installation and operating costs in a wind array. The goal is to maximize

the power per unit cost. In this study, a comprehensive model will be developed in which

turbines’ positioning is optimized, evaluated and analyzed under simulated various real

wind farm circumstance. Factors affecting the optimization and evaluation such as wind

distribution resolution, wake model, wind farm site limitation and turbine type will be

analyzed to improve the reliability of the program.

The objective is to provide an offshore wind array optimization tool suitable for

characteristic of any offshore array site. It analyzes the performance of wind turbines within

19
the array for given external wind distributions and optimizes the array layout to maximize

array power production per unit cost. It is expected to find some patterns of turbine

placement that can obtain more power on various wind farm sites.

20
Chapter 4 Wake models

As wakes develop downstream, they interact with atmospheric boundary layer as

well as with other wakes and are also affected by variable surface terrain. The extent of the

wake and the way in which the wake evolves is dependent on a number of factors. These

include the wind conditions (speed and turbulence intensity) entering the wind field (the so-

called inlet conditions), the terrain, the interaction of the wake with the atmospheric

boundary layer [15, 66], the turbulence developed in the wake (wake-added turbulence)

and the characteristics of the turbine (size, blade design, hub height, etc.).

Figure 4.1 Wake profile downstream a turbine

As wind passes through the region swept by the blades, over the nacelle and past

the tower, the energy extracted by the wind turbine reduces the wind speed. Figure 4.1

shows an idealization of the turbine wake. The near wake behavior is different from the far

wake region.

21
The wind speed measured immediately downstream of a turbine is significantly

lower than the free stream velocity. At near hub-height, the air in the wake is more

turbulent. This wake is characterized by the relative velocity deficit [26]

𝑈0 − 𝑈(𝑥)
𝑈′ = 4.1
𝑈0

where 𝑈(𝑥) is the velocity in the wake at a distance x downstream from the originating

turbine and 𝑈0 is the free-stream velocity. The wake velocity 𝑈(𝑥) and 𝑈0 are obtained

from meteorological data measured at fixed points in and around the wind field. Another

significant characteristic of the wake is its turbulence intensity [66]

𝜎𝑈
𝐼𝑤 = 4.2
𝑈�

� is the mean wind speed (in the wake) and U is the standard deviation of the wind
where 𝑈

speed in the wake. Turbulence intensity will affect wind turbine performance and different

𝐼𝑤 can result in different wind turbine power curve characteristics for the same wind speed

[66]. Compressibility as wind moves through the plane of the rotor is negligible and the

velocity reduction across the rotor plane results in a downstream conically expand wake.

The wake continues to expand downstream and, for flat terrain, interacts with the ground

when the wake radius equals the turbine hub height. Frandsen et al. [15] reported that this

occurs at a typical distance of ten rotor diameters downstream. Turbulent diffusion of

momentum from the free stream occurs due to the initially large velocity gradient and as

the air moves downstream the velocity deficit diminishes.

22
Wind turbine wakes have been studied for two decades and various models have

been developed. These models can be divided into two main categories, namely, kinematic

wake models (analytical wake models) and field wake models (computational or numerical

wake models). A kinematic wake model characterizes the velocity in a wake by a set of

analytical expressions whereas in field wake models, fluid flow equations, whether

simplified or not, must be solved to obtain the wake velocity field.

In what follows past work on wake models relevant to wind turbines is discussed.

The models range from simple linear wake models that simply account for downstream

attenuation of wind speed in the near wake region to reattainment of the free stream

speed in the far wake, to more sophisticated computational fluid dynamics based models

that can account for wake turbulence and in some cases uneven topography.

4.1 Kinematic Wake Models

Kinetic wake models are developed from the momentum equation to model the

velocity deficit in the wake behind a turbine. The wake descriptions usually do not consider

the near wake region (less than two turbine diameters distance downwind a turbine). They

also do not account for the change in turbulence intensity in the wake. Thus they have to be

combined with a turbulence model if it is required to account for values of the turbulence

intensity throughout the wind farm. Kinetic wake models are simple and computationally

economic in that they can be easily implemented within large scale for calculations of large

wind farm performance.

23
4.1.1 Jensen wake model

The so-called Jensen model is one of the oldest and simplest wake models

and developed by N.O. Jensen [26]. It has been used in several studies that employ a

wake model in algorithms that attempt to optimize the cost per unit power by seeking the

optimizing placement of wind turbines within a given area. He treated the wake behind the

wind turbine as a turbulent wake which ignores the contribution of vortex shedding that is

significant only in the near wake region. The wake model is thus derived by conserving

momentum downstream of wind turbine The velocity in the wake is given as a function of

downstream distance and it is assumed that the wake expands linearly downstream of the

wind turbine. The width of the wake is given by

𝐷𝑤 = 𝐷𝑟 (1 + 2𝑘𝑠) 4.3

and the velocity in the wake u is given by

1 − �1 − 𝐶𝑇
𝑢 = 𝑈0 [1 − ] 4.4
(1 + 2𝑘𝑠)2

where 𝐶𝑇 is turbine thrust coefficient, k is Wake Decay Constant and s=x/D is relative

downstream distance. The value of k is generally taken to be 0.075 for land cases, and 0.05

is recommended for offshore cases [41]. 𝐷𝑟 is “expanded” downstream rotor diameter of

turbine diameter D of the form

1−𝑎
𝐷𝑟 = 𝐷� 4.5
1 − 2𝑎

Katic et al. [40] expanded the previous work by Jensen, describes the wake velocity

profile as constant or ‘top hat’ profile which is justified by the fact that the purpose of the

24
model is to estimate the energy content within the wind field as seen by downwind turbines

rather than accurately describing the spatial variation velocity field. Notably, the Katic

model is currently the basis of the wake model used in WAsP. WAsP is a wind climate and

turbine power prediction software developed by Risø. It is also included in Garrad Hassan

WindFarmer and WindPRO [59-61].

4.1.2 Larsen model

The model developed by G.C. Larsen [42], also known as the EWTS-II model (used in

WindPRO), is based on the Prandtl turbulent boundary layer equations. A self-similar

velocity profile is assumed and Prandtl’s mixing length theory [43] is used to get a closed

form solution. The flow is further assumed to be incompressible, stationary and of no wind

shear, and thus the flow is axisymmetric.

Larsen showed both a first-order and a second-order approximate solution to the

boundary layer equations [67], of which the last one is capable of resolving the double dip

in the velocity deficit profile of the near wake. The following equations for the, rotor wake

radius 𝑅𝑤 and the axial velocity deficit in the wake (∆𝑈)1 are obtained

1
35 5 1 1
𝑅𝑤 (𝑥) = � � (3𝑐12 )5 (𝐶𝑇 𝐴(𝑥 + 𝑥0 ))3 4.6
2𝜋

3
𝑈∞ 1 3 1 35 10 1

(∆𝑈)1 (𝑥, 𝑟) = − (𝐶𝑇 𝐴(𝑥 + 𝑥0 )−2 )3 [𝑟 2 �3𝑐12 𝐶𝑇 𝐴(𝑥 + 𝑥0 )� 2 − � � (3𝑐12 )5 ]2 4.7
9 2𝜋

25
Two unknown constants, 𝑐1 is respectively related to the Prandtl mixing length and

𝑥0 is the position of the rotor which respect to the applied coordinate system. Following

relations are given by Larsen [42],

5 1
𝐷𝑒𝑓𝑓 2 105 −2 5
𝑐1 = � � � � (𝐶𝑇 𝐴 𝑥0 )−6 4.8
2 2𝜋

9.5𝐷
𝑥0 = 4.9
2𝑅
( 𝐷 9.5 )3 − 1
𝑒𝑓𝑓

Where the effective rotor diameter 𝐷𝑒𝑓𝑓 is calculated by

1 + �1 − 𝐶𝑇
𝐷𝑒𝑓𝑓 = 𝐷� 4.10
2�1 − 𝐶𝑇

And 𝑅9.5 is the wake radius at 9.5 rotor diameters downstream the turbine is taken

from [30]

𝑅9.5 = 0.5[𝑅𝑛𝑏 + min(𝐻, 𝑅𝑛𝑏 )] 4.11

With an empirical relation [42]

𝑅𝑛𝑏 = max(1.08𝐷, 1.08𝐷 + 21.7𝐷(𝐼𝑎 − 0.05)) 4.12

4.1.3 Frandsen analytical model

Frandsen’s model, also known as Riso’s analytical model [15], is designed to

estimate wake behavior across an entire wind farms rather than individual turbines. It is

based on conservation of the momentum deficit in the wake. The model distinguishes three

different wake regimes: a single wake regime, two neighboring interacted wake flow

regimes and a wake flow regime that is in balance with the atmospheric boundary layer.

26
In single wake case, the velocity deficit is determined as

1 𝐴0
𝑢= 𝑈0 ± �0.25 − 0.5𝑈02 𝐶𝑇 4.13
2 𝐴

where A0 is the turbine rotor swept area immediately downstream of the rotor, and A is the

area of the wake. The expansion of the area of the wake for the single wake case is given by

𝐷 1
= 3/2 4.14
𝐷0 (𝛽 + 𝛼𝑠𝑓 )1/3

where

1 1 + �1 − 𝐶𝑇
𝛽= ∙ 4.15
2 �1 − 𝐶𝑇

and α is a constant related to the thrust coefficient, which describes the initial wake

expansion and sf is the turbine spacing.

In the case of multiple wakes, the wakes are divided in several sections each having

a constant but different velocity deficit. In order to calculate the mean wind speed over the

rotor area, a semi-linear method is applied. For details see the paper by Rathmann et al.

[59].

4.2 Field Wake Models

Field models (also known as implicit models), in contrast to kinematic models,

calculate the flow at every position throughout the wake field, by solving the Reynolds-

averaged Navier-Stokes equations with a turbulence model for closure. The following are

two best known field models.

27
4.2.1 Ainslie wake model

The Ainslie wake model is a two-dimensional field model [57]. The model assumes

that the wake profile is axisymmetric and assumes a Gaussian profile at the two rotor

diameters downstream from the turbine. The flow is considered to be incompressible with

no external forces. Beyond the first two diameters, the gradients of mean quantities in the

axial direction are neglected as they will be much less than the gradients in the radial

direction [57]. The flow can then be described with a two-dimensional Reynolds equation in

the thin shear-layer approximation without viscous terms, as following

𝜕𝑈 𝜕𝑉 𝜖𝑣 𝜕𝑈 𝜕 2𝑈
𝑈 + 𝑉 = ( + 𝑟 2) 4.16
𝜕𝑥 𝜕𝑟 𝑟 𝜕𝑟 𝜕𝑟

𝜕𝑈 1 𝜕𝑉
= − �𝑟 + 𝑉� 4.17
𝜕𝑥 𝑟 𝜕𝑟

here 𝜖𝑣 is the eddy-viscosity and used for closure. These two equations are solved

numerically starting from the near wake with an empirical wake profile. The initial

centerline velocity (at x = 2D) deficit 𝐷𝑚𝑖 is given by equation

𝐼𝑎
𝐷𝑚𝑖 = 𝐶𝑇 − 0.05 − (16𝐶𝑇 − 0.5) 4.18
1000

where 𝐶𝑇 is the wind turbine’s thrust coefficient (a function of the upstream wind speed)

and 𝐼𝑎 is the free stream turbulence intensity.

The wake width 𝑏 which increases with downstream distance is related to the thrust

coefficient 𝐶𝑇 and centerline velocity deficit 𝐷𝑀

3.56𝐶𝑇
𝑏=� 4.19
8𝐷𝑀 (1 − 0.5𝐷𝑀 )

28
Given a wake width 𝑏 and a Gaussian shape by the wind velocity in the wake U is of

the form

𝑈 2
1− = 𝐷𝑀 𝑒 −3.56(𝑟/𝑏) 4.20
𝑈0

here 𝑈0 is the free stream velocity and 𝑟 is the distance from the centerline.

The effects of turbulence and turbine operation (described in 𝐶𝑇 ) are included in the

model. A lower turbulence intensity leads to a slower wake recovery. Higher 𝐶𝑇 leads to a

stronger near wake. Software packages that apply the Ainslie wake model include WindPRO,

GH WindFarmer, FLaP.

The Ainslie model was adapted by Anderson [68] who developed a simplified

solution. He substituted equation (4.17) and equation (4.20) into equation (4.16) and

obtained

𝑑𝑈𝑐 16𝜖𝑣 (𝑈𝑐3 − 𝑈𝑐2 − 𝑈𝑐 + 1)


= 4.21
𝑑𝑥 𝑈𝑐 𝐶𝑡

here the normalized centerline velocity is

𝑈𝑐 = (1 − 𝐷𝑀 ) 4.22

The system (4.16) - (4.17) now reduces to a first order differential equation (4.21)

which can be solved efficiently by a simple numerical integration scheme. In our study, we

applied this simplifying method and decreased computing time two orders of magnitude.

4.2.2 Three-dimensional field model

4.2.2.1 ECN Wakefarm model

29
The ECN model is based on the UPMWAKE model developed by Crespo et al. [69].

The model is a 3D parabolized Navier-Stokes code for the far wake using a k-𝜖 turbulence

model. The near wake model is solved by momentum theory and some empirical

corrections. The increase of momentum at infinity is subtracted from the undisturbed wind

profile. The near wake profile is

𝑢(𝑧) = (1 − 2𝑎)𝑈0 (𝑧) 4.23

where a is designated to the axial induction factor in the rotor plane. Similar to Jensen wake

model, the velocity decrease is assumed to be constant with the “expanded” rotor diameter

𝐷𝑟

1−𝑎
𝐷𝑟 = 𝐷� 4.24
1 − 2𝑎

The near-wake velocity deficit is defined as Gaussian in shape initial at 2.25D


2 2
𝑈𝑖𝑛𝑖𝑡𝑖𝑎𝑙 (𝑧) = 1.3(1 − �1 − 𝐶𝑇 )𝑈0 (𝑒 −0.5(𝑦/𝑟𝜎𝑦 ) 𝑒 −0.5((𝑧−𝐻)/𝑟𝜎𝑧 ) 4.25

where the factors 𝜎𝑧 and 𝜎𝑦 are defined as

𝐷𝑟
𝜎𝑧,𝑦 = 4.26
2𝐷

4.2.2.2 Elliptic field models

Researchers at Robert Gordon University (RGU) have developed a fully elliptic

turbulent three-dimensional Navier-Stokes numerical solver with k-ε turbulence closure

during the ENDOW project [63] based on a previous axisymmetric model by Magnusson et

al. [64]. Initial data required to start the 3D-NS calculations are the velocity and turbulence

30
intensity profiles in the atmospheric boundary layer upstream of the rotor. The

computational domain includes the rotor of the wind turbine, which is approximated by

means of a semipermeable disk to simulate the pressure drop across a real rotor disk. The

computational time is much longer for this model compare to other model [70].

4.3 Wake added turbulence models

As mentioned earlier, some wake models have to be combined with turbulence

intensity when used for wake calculations. In our work, we used the Frandsen

model/IEC61400-1 standard [71] to calculated added turbulence in wakes.

The model assumes the following: if the distance separation between two wind

turbines in a wind farm is more than 10 rotor diameters, wake effects on turbulence

intensity can be neglected. If the separation is less than 10 rotor diameter, the wake effects

on the added turbulence have to be calculated following equations. Wake added turbulence

intensity, which Frandsen calls ‘effective turbulence’, is given by

1
𝑁 𝑚
𝐼𝑒𝑓𝑓 = �(1 − 𝑝𝑤 𝑁)𝐼𝑎𝑚 + 𝑝𝑤 𝑁 � 𝐼𝑇𝑚 (𝑠𝑖 )� 4.27
𝑖=1

where m is the Wohler curve exponent depending on the material of the structural

component under consideration

N is the number of neighboring turbines (nearest and next nearest neighbors)

𝑝𝑤 is the wake probability and is taken to be 0.04

𝑠𝑖 is the distance to the neighboring turbine i

31
𝐼𝑎 depends on the considered load case

and the maximum center wake turbulence intensity 𝐼𝑇 is given by

1
𝐼𝑇 = � + 𝐼𝑎2 4.28
1.5 + 0.3𝑠𝑖 �𝑈0

The wake of a turbine that is upstream another turbine from the point of view of the

turbine in consideration is not taken into account. So in a rectangular array configuration,

the possible number of nearest and next nearest neighboring turbines is 8, see also Figure

4.2.

Figure 4.2 Turbines taken into consideration when calculating added turbulence

Inside large wind farms, the wind turbines tend to generate their own ambient

turbulence. Thus, when (a) the number of wind turbines from the considered unit to the

32
“edge” of the wind farm is more than 5, or (b) the spacing in the rows perpendicular to the

predominant wind direction is less than 3 rotor diameters from each other, a different value

for ambient turbulence intensity, Ia′ , is calculated by

1
Ia′ = 2 + I2 + I �
��Iw a a 4.29
2
0.36
𝐼𝑤 = 4.30
𝑠𝑟 𝑠𝑓
1 + 0.2�
𝐶𝑇

The values sr and sf are the relative spacings in the rows and between the rows of
turbine in wind array

33
Chapter 5 Methodology

This chapter describes the Wind farm performance evaluation and layout

optimization model used in subsequent chapters. The motivation for developing the model

is to enable the characterization of the in-situ operating environment within an offshore

wind array to supply quantitative information necessary to determine optimal turbine array

layout and even guide wind plant control/operation and power production estimates. In

previous chapters, the development of model approach that can provide quantitative

characterization of the in-situ operating environment throughout the array with sufficient

resolution to capture the essential dynamics within the array to quantify performance have

been presented and discussed. In the chapter, the wind array configurations are simulated,

wind data or wind distribution is processed to describe the annual average wind condition

and wake models are used to evaluate wake losses. In addition, a new multiple wake model

and a wind farm roughness model were developed by measured data from Horns Rev wind

farm and Nysted wind farm. It only needs a simple computation scheme for any complex

multiple wakes case, but its result is promising. An optimization model is also created in a

Matlab code and applies Matlab “Global Optimization Toolbox” to optimized wind array

turbine placement layout.

5.1 Large wind array performance evaluation model (LWAP)

In this section, a new model for large wind array performance evaluation to evaluate

wake effects in large wind farms is developed. To distinguish it from other models it is

34
referred to as the LWAP or the LWAP model. The LWAP applies a single wake model such as

the Jensen wake model or the Ainslie wake mode, and accounts for multiple wake

interactions and the effect on the vertical wind profile in the atmosphere boundary layer by

the wind farm itself. The LWAP predicts wind speed deficits at each turbine and for specific

turbine power curves and assesses power for individual turbines and for the entire wind

farm. The calculation method converges within seconds for a large wind farm evaluation.

The calculation scheme of the model comprises the following steps:

• Apply wind data (speed, direction and turbulence intensity) over the wind

farm site.

• Place the wind turbines in the wind farm and calculate the new free stream

wind velocity due to the increased surface roughness in the atmospheric boundary

layer caused by the presence of the wind farm.

• Input this new free stream velocity into a combination of a single wake

model and a multiple wake model to evaluate the wake deficits in the wind farm.

The single wake models are introduced in Chapter 4. Here the multiple wake model

and the model for the evaluation of the effect by the wind array on the atmospheric

boundary layer are developed. A Matlab code was developed for this simplified model that

is suitable for wind farm evaluation. In order to save computational time for the Ainslie

implicit model, a simplified solution for the model by Anderson [68] was used, and wind

speed deficits in the field of the single wake were calculated and saved as a data base for

using for wake deficits evaluation. The code requires much less computational time than

35
CFD based wind farm evaluation models, but the accuracy is not sacrificed. For comparative

purposes both the Jensen wake model and the Ainslie wake model were used as the single

wake model to predict wind speed deficits and turbine power deficits that were compared

to observations from the Horns Rev and Nysted wind farm.

5.1.1 Multiple wake model

A simplified multiple wake model is developed by considering the wake of only the

nearest upstream turbine and its operational condition (wind speed, thrust coefficient and

turbulent intensity). The following assumptions were made:

• The nearest upstream turbine always dominates in multiple wake effects.

• Wind turbine operating conditions and the extra flow momentum exchange

between multiple wakes and with the free stream will affect or contribute to

downstream wake recovery.

• The effect of overlapped wakes was a combination of momentum flow from a

single turbine wake and interaction with the free stream which can be attributed

to the higher momentum exchange. Thus the wind speed in the ith overlapped

wake, 𝑢𝑖 , is

𝑢𝑖 = 𝐶𝑢𝑖′ + (1 − 𝐶)𝑈0 5.1

where 𝑈0 is the free stream velocity, 𝑢𝑖′ is the velocity calculated in the wake of the

upstream turbine by a single wake model, and C is a combination coefficient having the

value from 0 to 1. The combination coefficient C is a function of turbine downstream

distance 𝑠𝑖 in that 𝐶 → 0 for small 𝑠𝑖 and 𝐶 → 1 for very large 𝑠𝑖 .

36
When the upstream turbine is in the wake of another turbine, the wind entering the

rotor area has a lower wind velocity than the free stream. Thus, if momentum exchange

with the atmospheric boundary layer is ignored the wind velocity will at most only recover

to the wind speed at the upstream turbine. However, because of interactions with the

atmospheric boundary layer, it should eventually recover to the free stream velocity as the

distance from the turbine gets large.

To determine the value for the combination coefficient C, the contribution of the

free stream velocity was assessed as shown in Figure 3. The wake in this case was produced

by a wind turbine in the free stream with a thrust coefficient, 𝐶𝑇 , equal to 1. This means

that no free stream momentum flow could pass through the turbine. The wake velocity,

(1 − 𝐶)𝑈0 , is then be calculated using equation (4.4) and gives

1
𝐶= 5.2
(1 + 2𝑘 ′ 𝑠)2

where 𝑘 ′ is the wake decay coefficient for this case. Its value should be less than 0.075

which is chosen for a normal wake, because there is a lower velocity gradient between the

inner wake and the free stream as shown in Figure 5.1. The wake decay constant, 𝑘 ′ , shown

in Figure 4, is a function of the ratio of upstream turbine velocity deficit to the free stream

velocity as following

𝑢𝑖−1 1.4 𝑢𝑖−1


𝑘 ′ = 0.8 �1 − � 0.8 < ≤1
𝑈0 𝑈0

𝑢𝑖−1
and 𝑘 ′ = 0.075 ≤ 0.8
𝑈0

where 𝑢𝑖−1 is wind speed at the location of upstream turbine.

37
Figure 5.3 shows the combination coefficient, C, calculated using equation (5.2) with

different 𝑘 ′ versus normalized downstream distance s. Figure 5.4 shows the wind speed

recovery calculated using the multiple wake model and the Jensen single wake model. The

wind velocity in the wake(s) caused by the wind turbine will finally recover to the free

stream velocity.

Figure 5.1 Overlapped wakes where 𝐬𝟏 =7D

38
0.2

0.18

0.16

0.14

0.12

0.1
k'

0.08

0.06

0.04

0.02

0
0 0.2 0.4 0.6 0.8 1
The ratio of wind velocity at upstream turbine to free stream velocity

Figure 5.2 Wake decay constant, 𝒌′ , as a function of upstream turbine wind speed deficit

Figure 5.2 shows the relationship between coefficient 𝑘 ′ and wind speed at upstream

turbine. When upstream turbine is in free stream, 𝑘 ′ is zero and will increase with

decreasing wind speed ratio to freestream which means higher wake effects on the turbine.

39
0.8

0.7

0.6

0.5

0.4
C

k'=0.030
0.3 k'=0.053
k'=0.075
0.2

0.1

0
0 10 20 30 40
Normalized downstream distance

Figure 5.3 Combination coefficient, C, as a function of normalized downstream distance

Figure 5.3 shows C calculated by equation (5.2), relating to upstream turbine wind

speed deficit and downstream distance. C is zero when downstream distance is zero,

because there is no space where the wake can interact with free stream. With downstream

distance increasing, C is increasing as the wake accumulates more free stream kinetic

energy. In addition, if the upstream turbine is under higher wake effects, it can be

considered to produce higher turbulence intensity. Then the increasing rate of C by the

downstream distance will be higher because higher turbulence intensity means more

energy exchanges.

40
0.9

0.8

0.7
Normalized wind velocity

0.6

0.5

1
Turbine in freestream
0.9 u1/Uo=0.81
u2/Uo=0.85
0.8
u3/Uo=0.9
0.7

0.6

0.5
0 10 20 30 40
Normalized downstream distance

Figure 5.4 Normalized wind velocity calculated using multiple wake model and the Jensen wake

single model. Blue: turbine in the free stream. Red: Second turbine with an incident wind velocity

of 0.81 of the free stream. Green: Second turbine with an incident wind velocity of 0.85 of the free

stream. Purple: Second turbine with an incident wind velocity of 0.9 of the free stream.

Figure 5.4 shows wind speed recovery calculated by our new multiple wake model.

Wind speed recovers slower than turbine wake in the free stream especially in the near

wake region. However, with downstream distance increasing (more than 3 to 4 D), the

recovery rate is similar to that from single wake model and will finally recover to free

stream wind speed.

41
5.1.2 Effect of the wind array on the atmospheric boundary layer

Investigations by Frandsen [15, 66] show that standard wake models under predict

wake effects. He states that the reason is that the effect a large wind farm has on the

atmospheric boundary layer is not taken into account. This effect was modeled as an

extended wind farm and resembles an increase in local surface roughness which results in a

change in vertical wind profile. According to this theory, the infinite wind farm equivalent

roughness 𝑧00 is calculated by

⎧ ⎫
𝜅
𝑧00 = ℎ𝐻 exp − 5.3
⎨ ⎬
�𝑐𝑡 + (𝜅/ln(ℎ𝐻/𝑧0 ))2
⎩ ⎭

where ℎ𝐻 is the hub height, 𝜅 is the von Karman constant, 𝑧0 is the local roughness without

wind farm, and 𝑐𝑡 is the distributed thrust coefficient defined by

𝜋
𝑐𝑡 = 𝐶 5.4
8𝑠𝑑 𝑠𝑐 𝑇

where 𝐶𝑇 is turbine thrust coefficient and 𝑠𝑑 and 𝑠𝑐 are the mean downwind and crosswind

spacing in rotor diameters. The wind farm roughness of 0.56 will be calculated for the Horns

Rev layout and 0.40 for the Nysted layout.

As the wind reaches the wind farm, an internal boundary layer (IBL) is modeled by

Sempreviva et al. [72] using an increase in roughness. The effect on the wind speed at the

hub height for the turbine within the array is estimated from meteorological theory under

the assumption that the upstream wind flow and the IBL flow are neutral logarithmic

42
profiles. In the model the free stream speed, 𝑈0 , for a known height, z, above the internal

boundary layer height, h, can be calculated by

𝑢∗0 𝑧
𝑈0 = 𝑙𝑛 � � 5.5
𝜅 𝑧0

Below the internal boundary layer, h, the free stream speed, 𝑈1 , is by

𝑢∗1 𝑧
𝑈1 = 𝑙𝑛 � � 5.6
𝜅 𝑧00

where 𝑢∗0 and 𝑢∗1 are the friction velocity for the free stream above and below the internal

boundary layer.

And combine equation (5.5) and (5.6), it is obtained

𝑈1 𝑢∗1 ln(𝑧/𝑧00 )
= ∙ 5.7
𝑈0 𝑢∗0 ln(𝑧/𝑧0 )

At the height h the wind speed above and below the internal boundary layer are the

same, hence

𝑢∗0 ln(ℎ/𝑧00 )
= 5.8
𝑢∗1 ln(ℎ/𝑧0 )

It follows then that when the height of the IBL grows higher than turbine hub height,

the new wind speed 𝑢 at turbine hub height can be calculated from equations (5.7) and (5.8)

and is

𝑈1 (ℎ𝐻 ) ln(ℎ/𝑧0 ) ln(ℎ𝐻 /𝑧00 )


= ∙ 5.9
𝑈0 (ℎ𝐻 ) ln(ℎ/𝑧00 ) ln(ℎ𝐻 /𝑧0 )

Frandsen et al. [15] developed a model of the growth of the internal boundary layer

for the overlapped wake case in the wind farm as following

𝜕ℎ 𝑐𝑚𝑤 𝑐𝑚𝑤
= 𝑐 → ℎ= 𝑐 (𝑥 − 𝑥𝑜 ) + ℎ0 5.10
𝜕𝑥 1 − 𝑐𝑚𝑤 𝑡 1 − 𝑐𝑚𝑤 𝑡

43
where 𝑐𝑚𝑤 is the relative flow speed in the wake, x is the downstream distance in the wind

farm and 𝑥𝑜 and ℎ0 are to be determined. Frandsen et al. [15] suggested that the value of

𝑐𝑚𝑤 could first approximated by the flow speed deficit in the infinitely large wind farm, 𝑐𝑤𝑓 .

Kristensen et al. [73] showed a relationship between the friction velocity before and after

the roughness change as following

𝑢∗1 𝑧00 0.07


≈� � 5.11
𝑢∗0 𝑧0

Hence the flow speed deficit in the infinitely wind farm, 𝑐𝑤𝑓 , can be calculated by

𝑈1 𝑧00 0.07 ln(ℎ𝐻/𝑧00 )


𝑐𝑤𝑓 = =� � ∙ 5.12
𝑈0 𝑧0 ln(ℎ𝐻/𝑧0 )

However, Frandsen’s model on the growth of internal boundary layer was developed

from a wind direction that is along turbine rows. In order to account for other orientations

this model is extended.

Based on observed data at Horns Rev [8, 60] the following assumptions were made

for our model of large wind farms:

• Offshore roughness height 𝑧0 is equal to 0.0001 without the wind farm [66].

• When a turbine is located downwind the array with a specified distance, 𝐷𝑠 ,

the height of the internal boundary layer reaches turbine hub height. This distance is

empirically assumed of 7 turbine rotor diameters, hence the integration constant 𝑥𝑜

in equation (20) is equal to 𝐷𝑠 = 7𝐷.

• Frandsen’s model on the growth rate of the internal boundary layer for the

wind parallel to wind turbine rows is extended for all other wind directions.

44
Based on these assumptions, the following empirical formula was developed for the

growth of the internal boundary layer for the wind farm

𝜕ℎ 𝑐𝑤𝑓 𝑐𝑤𝑓
= 𝑐 → ℎ(𝑥) = 𝑐 (𝑥 − 𝐷𝑠 ) + ℎ𝐻 5.13
𝜕𝑥 1 − 𝑐𝑤𝑓 𝑡 1 − 𝑐𝑤𝑓 𝑡

Figure 5.5 shows the calculated free stream speed at turbine hub height for the

Horns Rev farm layout.

1
Normalized freestream speed at turbine height

0.99

0.98

0.97

0.96

0.95

0.94

0.93

0.92
0 1000 2000 3000 4000 5000
Distance (m)

Figure 5.5 Predicted free stream wind speeds at turbine height for the Horns Rev wind farm layout

45
5.2 Wind array layout optimization model (WALOM)

In this section, the wind array layout optimization model is described. To distinguish

it from other models it is referred to as the WALOM or the WALOM model. It uses the LWAP

model to evaluate the turbine and wind farm power output. However, the multiple wake

model is simplified and effect on the wind profile in atmospheric boundary by the wind

farm is not considered in WALOM. Simulations of wind distribution, wind array layout and

turbine properties such as power curve were demonstrated. The Genetic Algorithm method

is also introduced and applied in the Matlab in which turbines’ positioning is optimized,

evaluated and analyzed under given simulated various wind farm circumstance.

5.2.1 Wind array configuration set up

The area occupied by the wind array is generally taken to be a rectangular area with

Cartesian coordinate system, see Figure 5.6. No matter whatever wind array site boundary

is, we can always find an appropriate rectangular to include all turbines’ locations. An x

coordinate and a y coordinate are given to each turbine to describe its location. Note that

all coordinates must be positive in our model for codes calculating wake effects.

For any given wind array, all wind turbine coordinates are then determined and

normalized ranging from 0 to 1. For example for Horns Rev array, turbine coordinates are

shown in Figure 5.7. In this way, we can create a general code which is suitable for wake

loss evaluation of any wind array configurations.

46
When evaluating wake effects, we first give each turbine a ranking arranged by

upwind location to downwind location order. For examples, when wind is from the north

direction, this ranking is ascending from north location turbine to south location turbine.

However, in East wind case, the ranking starts at east location turbine and ends at most

west location turbine, see Figures 5.8 and 5.9.

Figure 5.6 Wind farm boundary set up

47
Figure 5.7 Horns Rev array turbine layout coordinates

Figure 5.8 Horns Rev turbines ranking for North wind

48
Figure 5.9 Horns Rev turbines ranking for West wind

5.2.2 Wind data and distribution

The measurement and characterization of wind turbine wakes, while relatively

straight forward under certain controlled laboratory conditions, i.e., scaled wind turbine

models in wind tunnels or small turbines in large wind tunnels by Larwood [74], is more

complicated in the field because wakes are spatio-temporally variable phenomena shifting

with the direction of the wind and not amenable to meteorological measurements on a long

term basis without excessive instrumentation [21]. Erection and maintenance of

meteorological towers offshore is costly especially offshore and so wake measurements are

typically limited to measurements downstream of the prevailing wind direction. The Danish

49
wind turbine group at RISO has made the most advances in wake measurements. Three

meteorological towers erected in the vicinity of eleven turbines at the Vindeby offshore

farm have permitted simultaneous measurements of wind speed in the free stream and

wake for several wind directions by Frandsen et al. [20]. Measurements of velocity profiles

at different distances downstream from the turbines were also obtained by Barthelmie et al.

[23], using ship-borne Sonic Detection and Ranging (SODAR). Wake measurements are also

discussed by Barthelmie et al., in a recent paper [8]. Wake measurements have also been

obtained from satellite synthetic aperture radar (SAR) measurements by Christiansen et al.

[22].

5.2.2.1 Original wind data

In planning wind farms, shot-time wind data plays an important role in estimating

various engineering parameters, such as wake effect, power output, extreme wind load and

fatigue load. Raw data (including wind speed, wind direction etc.) will be tested for

validation and then be analyzed. Recording an average wind speed and wind direction of 10

minutes period is widely used. Table 5.1 is a recording sample 10 minutes average wind

data measured on Lake Eire by the US Army Corps of Engineers [75].

This first wind data input method is to direct applying wind data into our program.

However, tens of thousands data are created even only for one year period wind recording.

Evaluating wake effects with this large number of wind cases need a lot of CPU time

50
Table 5.1 Wind data example

Time Period Station number Wind direction Wind speed


2000.0312.1700 92069 -81.72 10.3
2000.0312.1710 92069 -81.72 5.4
2000.0312.1720 92069 -81.72 2.6
2000.0312.1730 92069 -81.72 3.3
2000.0312.1740 92069 -81.72 4.1
2000.0312.1750 92069 -81.72 4.6
2000.0312.1800 92069 -81.72 4
2000.0312.1810 92069 -81.72 7.8
2000.0312.1820 92069 -81.72 5.5
2000.0312.1830 92069 -81.72 4.8
2000.0312.1700 92070 -81.76 10.1
2000.0312.1710 92070 -81.76 11
2000.0312.1720 92070 -81.76 2.2
2000.0312.1730 92070 -81.76 7
2000.0312.1740 92070 -81.76 4.6
2000.0312.1750 92070 -81.76 5.4
2000.0312.1800 92070 -81.76 5.4
2000.0312.1810 92070 -81.76 4.1
2000.0312.1820 92070 -81.76 7.7
2000.0312.1830 92070 -81.76 4.2

51
5.2.2.2 Weibull wind distribution

The wind variation for a site can be statistically described using a Weibull

distribution [76]. Wind speed data in our study is discretized by 1m/s. Wind orientation

(angular) distributions is also discretized by different direction sectors. The probability

density function of the wind speed for a direction θj sector is

𝐾𝑗 𝑈0𝑗 𝐾 −1 −(𝑈 /𝐴 )𝐾𝑗


𝑓(𝜃𝑗 , 𝑈0𝑗 , 𝐴𝑐𝑗 , 𝐾𝑖 ) = ( ) 𝑗 𝑒 0𝑗 𝑐𝑗 5.14
𝐴𝑐𝑗 𝐴𝑐𝑗

where the direction-dependent Weibull parameters are 𝐾, the shape factor, 𝐴𝑐 the scale

factor.

Figure 5.10 Mean Wind Speed Frequency Distribution (Weibull shape factor, K, and scale factor,
𝐀 𝒄 , was 2.10 and 8.33 respectively)

52
Figure 5.10 shows an example of the Weibull wind distribution. The wind data is

from a 50m height anemometer. It shows a maximum frequency of 5.57% for the bin 6.5 to

7.0 m/s (14.5 to 15.7 mph). This is the modal point of the distribution. The average was

reported earlier to fall within the bin 7.0 to 7.5 m/s (15.7 to 16.8 mph) which occurs at a

frequency of 5.37%.

Figure 5.11 Wind Direction Rose

53
Figure 5.11 is a wind direction distribution rose for wind data from Lake Erie.

Prevailing winds come primarily from around southwest at nearly 45%. Southwest is the

most prevalent sector of the 16 with a frequency of 10.49%. Like the overall wind speed

frequency distribution, a direction wind rose cannot tell the whole story as about the wind

speed distribution. For each wind direction, Weibull factors and probabilities are calculated

from wind data and shown in Table 5.2.

Table 5.2 Weibull factors for different wind directions [77].

Sector 𝐀𝒄 K %

Mean 10.71 2.33 100.0

N 8.65 2.11 5.1

NNE 8.86 2.05 4.3

ENE 8.15 2.35 4.4

E 9.98 2.55 6.6

ESE 11.35 2.81 8.9

SSE 10.96 2.74 6.5

S 11.28 2.63 8.7

SSW 11.50 2.40 11.5

WSW 11.08 2.23 12.1

W 10.94 2.28 11.1

WNW 11.27 2.29 11.4

NNW 10.55 2.28 9.6

54
5.2.3 Cost function

The cost function is needed when optimizing wind array layout problem for a given

site information while number of wind turbine is also not determined. A cost function

related to number of wind turbines was chosen that was also used in previous studies [27-

31]. The total cost is only dependent on the number of wind turbines, N, installed in the

wind farm. The non-dimensional cost of wind farm is

2 1 2
𝑐𝑜𝑠𝑡 = 𝑁( + 𝑒 −0.00174𝑁 ) 5.15
3 3

The cost function is based on that a maximum discount of 1/3 is available when large

number wind turbines are purchased. Figure 5.12 shows the total cost of wind farm based

on number of turbines.

Figure 5.12 Cost of wind farm vs. number of turbines

55
5.2.4 Turbine power

Wind turbine extract kinetic energy from the wind. According to wind turbine

momentum theory, the available power from the wind crossing a wind turbine for a wind

speed, 𝑢, is

1
𝑃𝑎𝑣𝑎𝑖𝑙 = 𝜌𝐴𝑢3 5.16
2

Where A is turbine swept area and 𝜌 is air density

The turbine power output is by equation

𝑃 = 2𝜌𝐴𝑢3 𝑎(1 − 𝑎)2 5.17

Where 𝑎 is the induction factor which is relate to turbine thrust coefficient 𝐶𝑇

𝐶𝑇 = 4𝑎(1 − 𝑎) 5.18

Figure 5.13 shows power curve and correlated thrust coefficient for turbines

installed on Horns Rev wind farm. Note that turbine cut in speed is 4m/s and cut out speed

is 25m/s.

When determining a turbine power output in a wind array, first effect on

atmospheric boundary layer by existing wind array will be evaluated and an affected new

free stream field at the turbine hub height will be calculated. Then wind turbines are

checked in the upstream to downstream order that whether they are operating in the wake

of any other wind turbine. If this is not the case, then power is calculated using the

atmospheric boundary layer considered free stream velocity. Otherwise wind velocity at the

point where the wind turbine placed is determined by applying wake model discussed

earlier.

56
Figure 5.13 Vestas V80 2MW turbine power output and thrust coefficient vs. wind speed [78]

5.2.5 Wake effect evaluation

In WALOM model, the LWAP model is used to evaluate wake effects and predict turbine

power output and the whole wind farm performance with a given array configuration. However the

LWAP model used in WALOM was modified: (1) the multiple wake model was simplified that only

the wake by nearest upstream turbine were calculated; (2) the effect on the wind profile in

atmospheric boundary layer by wind farm itself is not considered.

57
5.2.6 Genetic Algorithm (GA)

The so called genetic algorithm has been a popular approach so far although it is not

the only option. The genetic algorithm is a stochastic global search method that evolves

transformations of a coded configuration [79, 80]. Genetic information is initially stored in a

“chromosomal” string that represents say, two individuals, and is used to create the genetic

code of a new individual with its own code. During the reproduction phase, each individual

is assigned a fitness value derived from its raw performance measure given by the objective

function. This value is used in the selection to bias towards more fit individuals. Genetic

operators manipulate the “genes” of the chromosomes between parent pairs by crossover

and mutation, producing child generation. After recombination and mutation, the individual

strings of new generation are then decoded, assigned with fitness values, and then

compared. The process continues through subsequent generation. Figure 5.14 shows a

simple GA operation in the optimization of wind array problems.

A wind farm layout evaluation code is developed in MATLAB which calculates the

power produced and the cost of a wind farm. The code is coupled with the genetic

algorithm solver (referred as ‘ga’ solver), available in MATLAB’s genetic algorithm toolbox

for optimization process. In our code, wind turbine position coordinates was developed. The

number of turbines N must be fixed first, and there are 2N variables represent X and Y

coordinates. In this algorithm, turbine can be placed any location compare to center of cells

in previous one. Information of turbines’ location coordinates is initially stored in

“chromosomal” strings that represent individuals, and is used to create the genetic code of

a new individual with their own codes. For given number of turbines, the placement of

58
these turbines is described in the genetic string. Evolution and adaptation of the individual

strings representing turbines’ placement will be applied and guarantee “best placement

individuals” with higher power output for a given wind distribution. In GA judgment of

better turbine placement is by the objective function, also referred to as the fitness value

𝑐𝑜𝑠𝑡
𝑂𝑏𝑗𝑒𝑐𝑡𝑖𝑣𝑒 = 5.19
𝑃𝑡𝑜𝑡𝑎𝑙

The flowchart in Figure 5.15 demonstrates the process through of the code and the

‘ga’ solver operates to find an optimal solution. The optimization process starts with the

initialization in the genetic algorithm ‘ga’ solver. In the initialization step, following

parameters are specified.

• Number of variables: The number of variables is twice the number of wind

turbines because two variables are required to specify the position of a wind

turbine in a two dimensional coordinates.

• Population size: The population size is the total number of individual solutions

that represent wind farm layouts in one evolution generation.

• Constraints: The constraints in the ‘ga’ solver are specified as bounds i.e., lower

and upper limits for the variables. The area size of the wind farm is specified in

constraints so that wind turbines cannot be placed outside the wind farm region.

• Optimization criteria: The optimization criteria are referred to as stopping

criteria in the ‘ga’ solver. It include maximum number of iterations which is also

referred as generations, stall generations (i.e., if average change in objective

59
function value over stall generations is less than function tolerance than

algorithm stops) and function tolerance.

After the initialization process, random set of solutions is created taking into account

the constraints. All the solutions created are analyzed by the wind farm layout evaluation

model. Estimated power production and the cost of the wind farm are calculated and

objective function value (cost per unit power) of each solution is returned to ‘ga’ solver.

In the next step, the optimization criteria are checked if they are satisfied or not.

When the optimization criteria are not met, all the solutions are ranked based on their

objective function values. A solution with small objective function value is better as its cost

per unit power is smaller and is placed before other solutions with larger objective function

value.

After ranking is completed, a number of solutions are selected and some new

solutions are created (reproduced). This selection of solutions is affected by the ranking

done in previous step and a solution with good ranking has a better chance of being

selected. New solutions are created while some solutions are copied from original set of

solutions to the new set of solutions. These selected a few solutions are one of the best in

terms of the ranking and are called elite count.

The last step before new set of generations (new population) is ready is called

Mutation. In this step some random changes are made in a few solutions. This step is very

important as it helps in maintaining diversity in the solution set. This new solution set is

analyzed by the wind farm performance evaluation code and this iterative procedure

continues until one of the optimization criteria is satisfied.

60
With application of a large number of evolutionary iterations, the sequential

placement configurations of turbines are iterated toward an optimal. Figure 5.15 shows a

flowchart of the optimization process of our work.

Figure 5.14 Genetic Algorithm process

61
Figure 5.15 Flowchart of Genetic Algorithm optimization process

62
Chapter 6 Results: wind farm layout evaluation model

Matlab codes for the LWAP model were developed to apply the Jensen wake and

Ainslie wake models to predict wake losses at the Horns Rev wind array and Nysted wind

arrays. Wind array configurations for wind farm and site information, wind distribution and

offshore conditions are taken into account. Power curves and parameters corresponding to

the Vestas V80-2.0 MW turbines located in Horns Rev and the Siemens 2.3 MW turbines

located in the Nysted Array are used in the calculations.

Results of wind speed deficits and turbine power deficits from multiple wind cases

are evaluated and compared to observation data by Rathmann et al. [59] and Barthelmie et

al. [8, 60]. Evaluations made using the model developed here are also compared to other

wind farm analyzing tools such as WAsP, WindFarmer and NTUA etc. Compared to other

models, our model applies simple computational schemes. For example, with a given wind

Weibull distribution on 72 wind direction sectors, an overall evaluation of wind array

efficiency will be processed in 2 seconds using the Jensen wake model and 10 seconds with

the Ainslie wake model.

6.1 Wind speed evaluations at Horns Rev when wind direction is along
turbine rows

In this case, two wind directions along wind turbine rows in the Horns Rev Array are

taken into consideration, shown in Figure 6.1. Wind turbine spacing is 7D in the 270° case

and 9.4D in the 222° case. Wind speed deficits at turbine locations in the row are evaluated

with two free stream speeds (8.5 m/s and 12m/s) and compared to observation data by

63
Rathmann et al. [60]. There are two groups of observation data, respectively from internal

row located in the interior of the wind farm and external row located at the edge of wind

farm. At these low to moderate wind speeds, the thrust coefficient is relatively high, and

thus the wake effects shown are likely to be severe.

Figure 6.1 Horns Rev layout: Case 1 of 270° and 7D spacing, Case 2 of 222° and 9.4 D spacing

64
1
Observation
0.95 Jensen model
Ainslie model
Wind velocity deficit

0.9

0.85

0.8

0.75

0.7
0 2 4 6 8 10 12
Turbine number

Figure 6.2 Horns Rev evaluation wind direction 270° and wind speed 8.5 m/s +/- 0.5 m/s

In Figure 6.2, evaluation of wind speed deficits at each turbine location from wind

direction of 270° and wind speed of 8.5 m/s is obtained and compared to observation data.

Results show that evaluation by Ainslie model in general under predicted wind speed deficit

by about 0.01 to 0.02 of free stream velocity. The Jensen model seems to have a better

prediction in this case.

65
1
Observed int. row
Observed ext. row
Jensen model
0.95
Ainslie model
Wind velocity deficit

0.9

0.85

0.8

0.75
0 2 4 6 8 10 12
Turbine number

Figure 6.3 Horns Rev evaluation wind direction 270° and wind speed 12 m/s +/- 0.5 m/s

In Figure 6.3, evaluation and observation data of wind speed deficits are from wind

case of 270° and 12m/s. Observations are from both internal turbine row and external

turbine row. Wind speed deficits at internal turbine are less than those at external turbine.

This may be because there is a more complex wake effect well inside the array, so that

turbulence intensity will be higher. This means that more energy is exchanged between the

turbine wakes and free stream wind. Clearly, the Jensen and Ainslie models generally over

predict the wind speed deficit in the wake. Predictions are better correlated to observations

66
from l turbine rows on the outside of the array. The Ainslie model appears to yield better

predictions than the Jensen model in this case.

0.95
Wind velocity deficit

0.9

0.85

Observed int. row


0.8 Observed ext. row
Jensen model
Ainslie model
0.75
0 1 2 3 4 5 6 7 8 9
Turbine number

Figure 6.4 Horns Rev evaluation wind direction 222° and wind speed 8.5 m/s +/- 0.5 m/s

In Figure 6.4, wind speed deficits are evaluated with a wind direction of 222° and

wind speed of 8.5 m/s. As discussed in Figure 6.3, observed wind speed deficits from

interior turbine rows are less than those from exterior rows. This may also because of

higher turbulence intensity inside the wind farm. Predictions by our model are not as good

as in the 270° wind direction case. Wind speed deficits are over predicted for the first three

turbines and under predicted for last two.

67
1

0.95

0.9
Wind velocity deficit

0.85

0.8

0.75 Observed int. row


Observed ext. row
0.7 Jensen model
Ainslie model

0.65
0 2 4 6 8 10
Turbine number

Figure 6.5 Horns Rev evaluation wind direction 222° and wind speed 12 m/s +/- 0.5 m/s

In Figure 6.5, it shows that predictions for wind speed of 12m/s on 222° wind

direction compared to observation.

To provide a quantitative evaluation of model performance versus the observations,

the mean absolute percentage error (MAPE) of the wind velocity was calculated using
𝑛
100% 𝑢𝑝𝑟𝑒𝑑𝑖𝑐𝑡𝑒𝑑 − 𝑢𝑜𝑏𝑠𝑒𝑟𝑣𝑒𝑑
MAPE = �� � 6.1
𝑛 𝑢𝑜𝑏𝑠𝑒𝑟𝑣𝑒𝑑
𝑖=1

where u is the wind velocity at each turbine (observed or predicted) and n is the number of

turbines.

68
The MAPE for the predictions by the LWAP using the Jensen wake model and the

Ainslie wake model versus observation data of mean wind speed are calculated in Table 6.1.

It shows that observations by the LWAP using both single wake models have MAPE less than

2.2%. Predictions from exterior turbines are better correlated to observation data (i.e.,

exhibit lower MAPE) than interior turbines. In addition, predictions for 270° wind direction

cases have lower MAPE than 222° wind direction cases.

Table 6.1 MAPE of the computed normalized wind velocity using observed data for two
wind directions and two wind speeds reported in [60] at Horns Rev.

Wind case LWAP – Jensen (%) LWAP – Ainslie (%)


Interior row Exterior row Interior row Exterior row
270° and 8.5 m/s 0.99 1.68
270° and 12 m/s 3.09 1.73 1.48 0.40
222° and 8.5 m/s 2.83 1.07 2.46 1.51
222° and 12 m/s 1.76 0.98 2.44 1.28
Average 2.17 1.19 2.01 1.21

6.2 Turbine power evaluations at Horns Rev: wind direction parallel to


turbine rows

Wind turbine power can be predicted by evaluating the wind velocity at all turbine

locations within the array and then calculating the predicted output power using a wind

turbine power curve [32]. The curve shown in Figure 6.6 illustrates the one used in the

model and represents a Vestas-V80 2 MW wind turbine in Horns Rev. Two wind direction

cases (the 270° case and 7D spacing, the 222° case and 9.4D spacing) along wind turbine

rows in the Horns Rev Array were examined.

69
6
x 10
2.5

Turbine power (w) 2

1.5

0.5

0
0 5 10 15 20 25 30
Wind velocity (m/s)

Figure 6.6 Power curve for the turbine at Horns Rev

0.95

0.9
Observation
0.85 Jensen model
Ainslie model
Normalized power

0.8

0.75

0.7

0.65

0.6

0.55

0.5
0 1 2 3 4 5 6 7 8 9
Turbine number

Figure 6.7 Turbines power at Case 1 for wind speed at 8m/s and direction 270° at Horns Rev

70
1

0.95

0.9
Observation
0.85 Jensen model
Normalized power

Ainslie model
0.8

0.75

0.7

0.65

0.6

0.55

0.5
0 1 2 3 4 5 6 7 8 9
Turbine number

Figure 6.8 Turbines power at Case 1 for wind speed at 10m/s and direction 270° at Horns Rev

0.95

0.9 Observation
Jensen model
Normalized power

0.85 Ainslie model

0.8

0.75

0.7

0.65

0.6

0.55
0 1 2 3 4 5 6 7 8
Turbine number

Figure 6.9 Turbines power at Case 2 for wind speed at 8m/s and direction 222° at Horns Rev

71
Figures 6.7 and 6.8 show results of wind turbine power calculated for a wind

direction of 270° and wind speeds of 8 m/s and 10 m/s. The second case is the wind

direction of 222° and a wind speed of 8 m/s (chosen because there is a real data set to

compare with). Results are shown in Figure 6.9. In this case, Barthelmie et al. [8] reported

data of the first 5 turbines because there was a large uncertainty in the measurements due

to small number of observations.

The mean absolute percentage error of the normalized turbine power was

calculated for each case using


𝑛
100% 𝑃𝑝𝑟𝑒𝑑𝑖𝑐𝑡𝑒𝑑 − 𝑃𝑜𝑏𝑠𝑒𝑟𝑣𝑒𝑑
MAPE = �� � 6.2
𝑛 𝑃𝑜𝑏𝑠𝑒𝑟𝑣𝑒𝑑
𝑖=1

where P is the normalized power to turbine power at the free stream at each turbine

(observed or predicted), and n is the number of turbines.

The MAPE for the turbine power are shown in Table 2. Predictions using the Ainslie

single wake model have a higher MAPE in 270° wind direction cases than those using Jensen

single wake model.

Table 6.2 MAPE of the computed normalized turbine power using computed predictions
and actual observations at Horns Rev [58]: wind direction parallel to turbine rows.

Wind LWAP – Jensen (%) LWAP – Ainslie (%)


270° and 8m/s 3.67 6.59
270° and 10m/s 4.11 5.61
222° and 8m/s 2.00 0.86
Average 3.26 4.35

72
6.3 Power output predictions for turbines in the row at the Horns Rev
and Nysted for a representative wind speed and variable wind
directions

Detailed case studies of turbine power losses due to wakes at the Horns Rev and

Nysted wind farms were analyzed. The major difference between the two wind farms is the

turbine spacing, with 7 × 7D at Horns Rev and 5.8 × 10.5D at the Nysted. The average power

at each turbine was calculated for seven wind directions: a wind direction where the flow is

down an exact row (ER) including observations within ± 2.5° (270° ± 2.5° at Horns Rev, 278°

± 2.5° at Nysted), and six directions of ER + 5°, + 10°, + 15°, - 5°, - 10° and - 15°, as shown in

Figures 6.10 and 6.11. The observed data were reported by Barthelmie et al. [60] and

results by wind farm layout evaluation models such as WAsP, WindFarm, WakeFarm and

NTUA were also reported and analyzed. As shown in Figure 6.12 and 6.13, results from

Horns Rev and Nysted illustrated an acceptable agreement between predictions and

observations in most directions except 255° in Horns Rev and ER ± 5° in both Horns Rev and

Nysted arrays. It was reported that in 255° wind direction at Horns Rev the asymmetry in

the observations may not reflect real case and it could be a data issue due to insufficient

observations [60].

73
Figure 6.10 Horns Rev Array. Exact Row (ER=270°) of turbines [60]

Figure 6.11 Nysted array. Exact Row (ER=278°) of turbines [60]

74
Figure 6.12 Normalized power at Horns Rev for the free stream wind speed of 8 ± 0.5 m/s:
comparison of models with observations

75
Figure 6.13 Normalized power at Nysted for the free stream wind speed of 8 ± 0.5 m/s:
comparison of models with observations

76
In order to compare predictions by the model to reported results of other wind farm

layout evaluation models, the root mean square deviation (RMSD) of the normalized

turbine power was calculated for each case using [60]

2
∑𝑛𝑖=1�𝑃𝑜𝑏𝑠𝑒𝑟𝑣𝑒𝑑 − 𝑃𝑝𝑟𝑒𝑑𝑖𝑐𝑡𝑒𝑑 )�
RMSD = � 6.3
𝑛

The RMSD of the turbine power predictions by models versus observation data of

the mean turbine power by various wind directions and wind speed of 8 m/s at Horns Rev

and Nysted wind farms are given in Tables 3 and 4. In general, predictions in by the LWAP

using either the Jensen and Ainslie single wake model have the same accuracy, with the

overall RMSE of 0.08 by using the Jensen model compare to 0.07 by the Ainslie model in

Horns Rev case. The overall RMSE for Nysted wind farm are 0.05 computed using the Jensen

model and 0.04 b using the Ainslie model. However, results computed using the Jensen

model show that there are large discrepancies between predictions and observations at

Horns Rev for a 265° and 275° wind direction cases and also for the Nysted 283° wind

direction case. This is likely due to the uniform wind velocity distribution in Jensen wake

model which tends to desensitize the sensitivity of this model to wind direction.

Compare to the RMSD by other wind farm layout evaluation models, the LWAP

performs better (i.e., exhibit lower RMSD) than WAsP. These predictions also have little

RMSD difference compare to CFD based models such as WindFarmer, WakeFarm and NTUA.

However, the run times for the LWAP applied to a full wind farm simulation like Horns Rev

wind farm is only a few seconds. This is significant less than reported time for running in the

CFD based WindFarmer, WakeFarm and NTUA [60].

77
Table 6.3 RMSD of the computed normalized power for various wind directions and a
wind speed of 8 m/s at Horns Rev wind farm [60]

Horns Rev: 8.0 ± 0.5 m/s


Direction(�) WindFarmer WakeFarm WAsP NTUA LWAP - LWAP -
Jensen Ainslie
255 0.15 0.17 0.07 0.12 0.17
260 0.05 0.07 0.17 0.11 0.08 0.10
265 0.05 0.03 0.03 0.10 0.04
270 0.04 0.08 0.06 0.08 0.05 0.02
275 0.04 0.05 0.04 0.14 0.09
280 0.03 0.03 0.11 0.03 0.04 0.04
285 0.05 0.03 0.10 0.04 0.02
All 0.07 0.08 0.12 0.07 0.08 0.07

Table 6.4 RMSD of computed normalized power for various wind directions and a wind
speed of 8 m/s at Nysted wind farm [60]

Nysted: 8.0 ± 0.5 m/s


Direction(�) WindFarmer WakeFarm WAsP LWAP - LWAP -
Jensen Ainslie
263 0.09 0.05 0.04 0.04
268 0.05 0.04 0.21 0.03 0.03
273 0.13 0.10 0.02 0.10
278 0.04 0.06 0.09 0.05 0.05
283 0.04 0.03 0.13 0.03
288 0.04 0.03 0.12 0.06 0.06
293 0.03 0.04 0.03 0.08
All 0.07 0.06 0.15 0.05 0.04

78
Chapter 7 Results: wind farm layout optimization

In this chapter, the wind array layout optimization model (WALOM) is applied with

Genetic Algorithm method to the wind farm layout optimization problem. Wind turbine

layout is optimized by a given site information and wind distribution, while wake effects are

minimized and therefore the expected power production is maximized.

Two approaches were implemented. One approach is an extension of the previous

work started by Mossetti et al. [25], in which a 2km × 2km square wind farm site area is

considered. Mossetti also assumed a simple wind distribution with a higher frequency for

some directions and applied the Jensen wake model in the calculation. In this work the

Mossetti approach was extended by developing a new turbine placement coordinates and

considering effect on the results by factors as turbine spacing limit, area size and wake

models. Results from this approach are shown in section 7.1.

Section 7.2 describes the second approach which is from the view point of the

existing or potential wind farm layout design. Unlike the approach discussed in section 7.1,

the wind data, site area, turbine properties etc., from a real wind farm site were used.

Weibull distributions were applied and turbulence intensities of wind were also factored

into the model. A real wind turbine power curve was used to convert wind speeds to power.

7.1 Extension of Mossetti’s approach

Three cases for an area of 2 km ×2 km square were investigated from section 7.1.1

to section 7.1.3. Case 1 is a simple example for a uniform wind direction with a free stream

velocity of 12 m/s, shows in Figure 7.1. There are two reasons for looking at this, one is to

79
compare and validate against previous work and two, understand the effects of parameter

variation in the model. For Case 2 the wind direction is variable and the free stream velocity

is constant at 12 m/s. There is an equal probability that wind blows from any direction. The

wind direction is discretized in 36 segments each measuring 10°. Case 3 is variable wind

direction and variable wind speed case. Figure 7.2 shows the wind distribution in this case.

Three wind speeds are possible, 17, 12 and 8 m/s. The probability is higher for wind

directions between 270o to 350o.

The optimization for wind Case 1 (uniform one speed wind) was also applied for

different wind turbine spacing limits (the smallest distance between two turbines) and area

sizes (2 km2, 4 km2, 8 km2 and 16 km2). Results are shown in sections 7.1.4 and 7.1.5.

In section 7.1.6, the optimization was applied to wind Case 2 (equal probability

variable wind direction with a constant wind speed) with a round shape wind farm site area.

The reason for studying on this is that it is expected to find some optimized layout pattern

for the turbine layout under isotropic wind distribution and isotropic site shape area.

In section 7.1.7, optimized turbines layouts using the Jensen and Ainslie wake

models were calculated and compared for the model three wind cases presented above.

7.1.1 Case 1: Constant unidirectional wind

The optimization was carried out for different numbers of turbines (N). For each N,

there is an optimal configuration. The space limit between two turbines is such that two

turbines should not occupy the same 100m × 100m square. Figure 7.3 shows optimal fitness

values for different N. The limiting value of fitness is the situation when there is no wake

80
effect for all turbines so that a maximum power is produced. It is noted that when N

increases, the difference between the optimal fitness value and the limiting value increases.

The optimal result is when N=48 and is compared with previous studies in Table 7. 1.

Figure 7.4 shows the computed optimal distributions of wind turbines from previous

studies and this study. All were configured under the same parameters except that in the

present study two turbines cannot occupy the same 100m × 100m square. The grid is

sufficiently refined following Mittal’s approach [29] that turbines can be placed anywhere

outside of the spacing limit.

In the Figure 7.4 (E), it shows that turbines trend to be positioned at upstream area

and downstream area. This is because in upstream area, there is no wake affect and

downstream area has better wake recovery. This is unlike the optimized layouts by other

works, shown in Figure 7.4. It may because of a different multiple wake model used in

WALOM. It is also noted that placement of diamond shape is preferred in the distribution as

it reduces the possibility that wind speeds will be lowered at the turbine because of the

velocity deficit in the wake.

Table 7.1 compares the number of turbines, the total power produced, fitness value

and efficiency for the results of previous studies and the present study. The optimal

placements of previous studies were also recomputed by our model. The power outputs of

previous configurations were reduced in the recomputed results.

81
Figure 7.1 Wind farm area

Figure 7.2 Wind distribution for case 3

82
Figure 7.3 Fitness value of different number of turbines for case 1.

Table 7.1 Results from previous study and current study: reported and recomputed

Mossetti et al. Grady et al.


Reported Recomputed Reported Recomputed
Number of 26 26 30 30
turbines
Total power 12352 12165 14310 13969
(km)
Fitness value 1.6197×10-3 1.6446×10-3 1.5436×10-3 1.5813×10-3
Efficiency (%) 91.645 89.785 92.015 89.344
Marmidis et al. Mittal et al. WALOM
Reported Recomputed Reported Recomputed Report
Number of 32 32 44 44 48
turbines
Total power 16395 8621 21843 21937 24529
(km)
Fitness value 1.4107×10- 2.6830×10-3 1.3660×10- 1.3602×10-3 1.3164×10-3
3 3

Efficiency (%) N/a 51.691 95.762 95.667 98.056

83
A B
↓ ↓

C D E
↓ ↓ ↓

Figure 7.4 Turbines placement of four studies for case 1: (a) Mossetti et al. [25] (b) Grady et al.
[27] (c) Marmidis et al. [28] (d) Mittal et al. [29] (e) WALOM

84
7.1.2 Case 2: Constant wind speed with an equal probability variable wind direction

In this case, several optimal layout configurations were obtained. Figure 7.5 shows

the fitness values for different number of turbines. When N=32, the fitness value reduces to

a minimum of 1.5749×10-3. The optimized placement of the turbines for this case and those

obtained by previous study are shown in Figure 7.6.

Figure 7.5 Fitness value of different number of turbines for case 2

85
A B

C D

Figure 7.6 Turbines placement for case 2: (a) Mossetti et al. [25] (b) Grady et al. [27] (c) Mittal et
al. [29] (d) WALOM

Table 7.2 is a comparison for the results of previous studies and the present study.

Even though the optimal configuration obtained does not have highest efficiency the fitness

value is lowest.

Table 7.2 Results from previous study and current study: reported and recomputed

Mossetti et al. Grady et al. WALOM


Reported Recomputed Reported Recomputed Report
Number of 19 19 39 39 32
turbines
Total power 9244 9099 17220 15728 14687
(km)
Fitness value 1.7371×10-3 1.7634×10-3 1.5666×10-3 1.7117×10-3 1.5749×10-3
Efficiency (%) 93.851 91.895 92.174 77.383 88.069

86
7.1.3 Case 3: Variable Wind Speed with Variable Wind Direction

In this case, the optimal configuration was obtained when N=35. Figure 7.7 shows

optimal fitness values for different number of turbines. The optimal wind farm layout

configuration of previous studies and that obtained using the model described earlier are

showed in Figure 7.8.

Figure 7.7 Fitness value of different number of turbines for case 3

In comparison to the previous optimal configurations, it has been shown that

optimal wind turbine location will tend to be positioned in the edges of the turbine array

area. Presumably this is an attempt to maximize the distance between turbines and to

87
eliminate the effects of wakes, like what had been discussed in section 7.1.1. It is worth

noting that, because the proximity exclusion rule has been relaxed in comparison to the

restrictions in previous models which did not allow turbines to occupy the same 200m ×

200m square, turbines at the edges cluster in a zigzag fashion along the edge direction. This

is to reduce wake effects when the wind blows along the edge of the square.

Table 7.3 compares the results obtained in previous studies and this study. It shows

that the efficiencies recomputed of previous configurations were reduced. The optimal

configuration by present study has a lower fitness value.

A B

C D

Figure 7.8 Turbines placement for case 3: (a) Mossetti et al. [25] (b) Grady et al. [27] (c) Mittal et
al. [29] (d) WALOM

88
Table 7.3 Results from previous study and current study: reported and recomputed

Mossetti et al. Grady et al. WALOM


Reported Recomputed Reported Recomputed Report
Number of 15 15 39 39 35
turbines
Total power 13460 13594 32038 29224 29478
(km)
Fitness value 9.941×10-4 9.843×10-4 8.031×10-4 9.212×10-4 8.385×10-4
Efficiency (%) 94.62 93.50 86.619 77.307 86.89

7.1.4 Different Spacing Limits for Unidirectional Wind

In this section, results were obtained by setting spacing limits of 50m × 50m, 100m ×

100m, 150m × 150m and 200m × 200m, for the case of unidirectional wind of the same 2km

by 2km square area. Four groups of data about fitness value and efficiency are compared.

Figure 7.9 shows the fitness values changes depending on the number of turbines and

Figure 7.10 shows the fitness values depending on the number of turbines for each spacing

limit.

The fitness value increases and the efficiency decreases faster for large spacing limit,

especially when there are more than 35 turbines.

89
Figure 7.9 Fitness values of different spacing limits

Figure 7.10 Efficiencies of different spacing limits

90
50m × 50m 100m × 100m

150m × 150m 200m × 200m

Figure 7.11 Optimal placements of four spacing limits for 40 turbines

Optimal placements of 40 turbines for the four spacing limits are plotted in Figure

7.11. More wind turbines are positioned in the center area of the square for large spacing

limits. It is probably because large spacing limit provides less placement options in upstream

and downstream areas.

91
7.1.5 Different Area Sizes for Unidirectional Wind

Optimization for area sizes of 2 km2, 4 km2, 8 km2 and 16 km2 were applied. The

spacing limit is 100 m ×100 m in the case. Figures 7.12 and 7.13 show the fitness values and

efficiencies of for each area size. Results indicate that the velocity deficit is much larger for

an area of 2 km2, but 35 turbines can still have a total efficiency of 95%. With increasing

area, the overall efficiency is reduced.

Figure 7.14 shows the optimal placement of 40 turbines for each area size. It is

noted that for all cases, the optimal positions of turbines tend to cluster on upstream and

downstream edges. This is consistent with the discussion for placement of wind turbine for

Case 1. It is also found that for the square area of 2 km2, the array efficiency drops rapidly

with increasing number of turbines for a turbine rotor diameter of 40 m.

92
Figure 7.12 Fitness values of different area sizes

Figure 7.13 Efficiencies of different area sizes

93
2 km2 4 km2

8 km2 16 km2

Figure 7.14 Optimal placement of four area sizes for 40 turbines

94
7.1.6 Constant Wind Speed with Variable Wind Direction and a circular site area

In this case, the wind distribution was the same as Case 2 described in section 7.1.2.

However, the optimization was applied to a circular wind farm area with a diameter of 2 km.

Figure 7.15 shows the fitness values for different number of turbines for this case. The

optimized placements for this case are shown in Figure 7.16.

-3
x 10
2

1.9 Limit value of fitness

1.8

1.7
Fitness value

1.6

1.5

1.4

1.3

1.2
0 5 10 15 20 25 30 35 40 45 50
Nuber of turbines

Figure 7.15 Fitness value as a function of turbine number

95
20 turbines 30 turbines
2000 2000

1500 1500

1000 1000

500 500

0 0
0 500 1000 1500 2000 0 500 1000 1500 2000
40 turbines 50 turbines
2000 2000

1500 1500

1000 1000

500 500

0 0
0 500 1000 1500 2000 0 500 1000 1500 2000

Figure 7.17 Optimal turbines layouts for a circular site area

Table 7.4 Turbines distributions in optimized layouts for isotropic wind and round area

Number of turbines Near area boundary Inside the area % of turbines near the
boundary
20 15 5 75
30 21 9 70
40 27 13 68
50 32 18 64

96
Figure 7.17 shows that wind turbines tend to be optimally positioned at the

boundary of the wind array area. This is similar to results obtained for variable wind

direction and square wind arrays presented earlier. Table 7.4 shows the number of turbines

near the site boundary versus the number of turbines inside the area. It shows that optimal

turbine locations are located at the site boundaries. With an increasing number of turbines,

more turbines will be optimally positioned within the interior area of the site. This can be

explained simply as an attempt to maximize the distance between turbines and thus

minimizing the wake effects of wakes.

7.1.7 Comparison of optimized layouts using Jensen wake model and Ainslie wake
model

In this section, optimized turbines layouts using Jensen wake model and Ainslie wake

model were obtained for 48 turbines with uniform wind (Case 1) and 32 turbines with

variable wind direction (Case 2) and a square area. Results are shown in Figures 7.18 and

7.19.

Figure 7.18 shows that optimal locations obtained using the Ainslie wake model

differ from those obtained using the Jensen wake model and tend to be located near the

upstream area and downstream boundaries of the array area. This is likely due to

assumption of a Gaussian wind velocity profile in the Ainslie wake model. In comparison to

the uniform wind velocity used in Jensen wake model, if a turbine is located at the edge of a

wake near another turbine, the wake deficit would be much less for the Ainslie model. Thus

the ‘proximity constraint’ that requires some distance between neighboring turbines due to

97
wake effects is relaxed somewhat for the Ainslie model. . In addition, turbine spacing in the

windward direction is greater than for lateral or crosswind directions.

Figures 7.19 and 7.20 show optimized layouts for Case 2 and Case 3 wind

distributions using both wake models. A major difference between optimized layouts

obtained using the Ainslie wake model to those obtained using the Jensen wake model is

that turbines show less of a tendency to be positioned near the boundary of the area. This

is especially evident in Case 3, where more turbine spacing in the prevailing wind direction

than crosswind direction was also observed.

Jensen wake model Ainslie wake model


2000 2000

1800 1800

1600 1600

1400 1400

1200 1200

1000 1000

800 800

600 600

400 400

200 200

0 0
0 200 400 600 800 1000 1200 1400 1600 1800 2000 0 200 400 600 800 1000 1200 1400 1600 1800 2000

Figure 7.18 Optimal turbines layouts for 48 turbines with Case 1 uniform one direction (from
North to South) wind using Jensen wake model and Ainslie wake model.

98
Jensen wake model Ainslie wake model
2000 2000

1800 1800

1600 1600

1400 1400

1200 1200

1000 1000

800 800

600 600

400 400

200 200

0 0
0 200 400 600 800 1000 1200 1400 1600 1800 2000 0 200 400 600 800 1000 1200 1400 1600 1800 2000

Figure 7.19 Optimal turbines layouts for 48 turbines with Case 2 for variable wind direction and a
constant wind speed using Jensen wake model and Ainslie wake model.

Jensen wake model Ainslie wake model


2000 2000

1800 1800

1600 1600

1400 1400

1200 1200

1000 1000

800 800

600 600

400 400

200 200

0 0
0 200 400 600 800 1000 1200 1400 1600 1800 2000 0 200 400 600 800 1000 1200 1400 1600 1800 2000

Figure 7.20 Optimal turbines layouts for 48 turbines with Case 3 for variable wind direction using
the Jensen wake Ainslie wake models.

99
7.2 Horns Rev wind farm layout optimization

Motivated by the results presented in the previous section an optimization approach

is applied the spatial configurations of wind turbines an actual offshore wind arrays. This

will give us a chance to compare a predicted optimal configuration to the actual

performance of that wind array. For the study the power curve, thrust coefficients and size

characteristics of a Vestas V80 wind turbine (This is the actual wind turbine installed at the

Horns Rev)e employed. The power and thrust coefficients are given in Table 7.4 [78]. The

direction-dependent Weibull parameters 𝐾, the shape factor, A𝑐 the scale factor and

probabilities as a function of wind orientation sector are shown in Table 7.5 [77]. The

average turbulence intensity in the Horns Rev for different wind speeds was computed by

Nino et al. [81] and is shown in Table 7.6.

To evaluate the sensitivity of angular discretization of the wind directions three

cases were considered: 12, 36 and 72 possible directions. Optimized turbines layout results

applying these three cases were obtained and are presented in sections 7.2.1 to 7.2.3. In

section 7.2.4, optimized results obtained using Jensen wake model are discussed and

compared to that using Jensen wake model.

100
Table 7.4 Vestas V80 thrust coefficient and power as a function of wins speed

Wind speed Power (kW) Thrust coefficient


4 66.6 0.818
5 154 0.806
6 282 0.804
7 460 0.805
8 696 0.806
9 996 0.807
10 1341 0.793
11 1661 0.739
12 1866 0.709
13 1958 0.409
14 1988 0.314
15 1997 0.249
16 1999 0.202
17 2000 0.167
18 2000 0.140
19 2000 0.119
20 2000 0.102
21 2000 0.088
22 2000 0.077
23 2000 0.067
24 2000 0.060
25 2000 0.053

101
Table 7.5 Wind distribution for Horns Rev

Sector A𝑐 K %
Mean 10.71 2.33 100.0
N 8.65 2.11 5.1
NNE 8.86 2.05 4.3
ENE 8.15 2.35 4.4
E 9.98 2.55 6.6
ESE 11.35 2.81 8.9
SSE 10.96 2.74 6.5
S 11.28 2.63 8.7
SSW 11.50 2.40 11.5
WSW 11.08 2.23 12.1
W 10.94 2.28 11.1
WNW 11.27 2.29 11.4
NNW 10.55 2.28 9.6

Table 7.6 Turbulence intensities for variable wind speeds

Speeds m/s I𝑎 % Speeds m/s I𝑎 %


4 15.0 15 8.3
5 10.1 16 8.5
6 9.7 17 9.1
7 9.6 18 9.7
8 9.0 19 10.5
9 8.2 20 10.6
10 8.4 21 10.7
11 8.3 22 10.8
12 8.1 23 10.9
13 8.6 24 11.0
14 8.4 25 11.0

102
7.2.1 Weibull data for 12 wind direction sectors

In this section, the effect of different angular (sector) discretizations was

investigated. The Weibull data is used as described above. Each sector is 30°. The direction

distribution is shown in Figure 7.21.

The optimized wind farm layout of 80 turbines is shown in Figure 7.22. The average

power output of the wind array is 80.4 MW with an improvement of 8.8% over the

simulated Horns Rev. The optimized spatial configuration captured an additional 57 GWh.

Figure 7.23 shows the total annual energy produced in GWh of all turbines for each

sector, for Horns Rev and the optimized configurations. It is noted that the most

improvement occurs when wind flows parallel to Horns Rev turbine array direction. This is

the case when wake effects are considerable.

Figure 7.21 Wind direction distribution for 12 direction sectors

103
Figure 7.22 Optimized layout for case 1

Figure 7.23 Annual wind power rose with 12 direction sectors

104
7.2.2 Wind data for 16 wind direction sectors

Optimization results may be affected by the angular discretization of the wind

distribution. The 12 sector resolution of the wind direction data may not be sufficient so the

consequences of adding more angular resolution into the description of the wind direction

input data is investigated. To test that whether a significant difference in placement

configurations will arise for different angular discretizations, some test simulations were

carried out. An actual Horns Rev wind distribution for 16 sectors was prepared and input

into the simulation model. The angle between two sector sides is 22.5°. For the 16-sector

case, the wind speed probability was calculated for each sector using the average Weibull

parameter for the Horns Rev wind field. Figure 7.24 shows the probability distribution for a

16 sector field. Figure 7.25 shows optimized result for 12 direction sectors. Obviously, the

optimized placement for 12 sectors is not the “optimal” configuration.

For most wind direction sectors, the actual rectangular array layout of Horn Rev has

better performance. In addition, turbines in the rectangular layout can obtain 641 GWh of

annual power, more than the 636 GWh calculated for the optimization layout using 16 wind

direction sectors. This demonstrates that the degree of resolution of the angular

discretization of the input wind data can significantly influence the optimization results.

105
Figure 7.24 Wind direction discretization for 16 direction sectors

Figure 7.25 Annual power comparison of 16 direction sectors

106
7.2.3 Wind data for 72 wind direction sectors

To address the problem discussed above, more wind direction data with higher

discretized resolution is needed (if the spirit of ‘grid-refinement’ to demonstrate grid

independence in standard CFD calculations is adopted). To illustrate this, a synthesized wind

data with 72 orientation sectors was applied. The optimal layout configuration for this wind

field was obtained and compared to the predicted annual power output for the Horns Rev

configuration subject to the same wind field. As expected, the optimized configuration

produces more power. The wind direction probability density is shown in Figure 7.26.

Figure 7.26 Wind direction discretization for 72 direction sectors

107
Figure 7.27 Optimized layout for 72 direction sectors

Figure 7.27 shows the optimized layout. For this idealized case with constant wind

speed, the improvement is of 5.4% over the computed power for the actual Horns Rev

layout but subject to the synthesized wind field– 668 GWh compare to 704 GWh. It appears

that, with more wind direction sectors, there are more instances where there is an

increased probability that a turbine lies within in the wake of proximal upstream turbines.

Note that, even for this idealized wind field, the improvement of efficiency by 5.4% is not

inconsiderable in terms of actual power in a large array.

108
Table 7.7 Layout performcance of optimization results

Wind direction Layout configuration Power Annual layout Capacity factor

12 direction Horns Rev 72.0 MW 631 GWh 0.45


discretization
Optimization 12 wind 80.4 MW 704 GWh 0.50
directions

Optimization 72 wind 77.8 MW 682 GWh 0.49


directions

72 direction Horns Rev 76.2 MW 668 GWh 0.47


discretization
Optimization 12 wind 66.3 MW 590 GWh 0.40
directions

Optimization 72 wind 80.4 MW 704 GWh 0.50


directions

Horns Rev data Report 68.5 MW 600 GWh 0.43

In Table 7.7, Horns Rev Array and two optimized array configurations from two wind

direction discretization cases (12 wind directions and 72 wind directions) were evaluated by

both wind discretization methods. It shows that the optimized configuration with 12

direction discretization produces less power than predicted for either the Horns Rev

configuration or the optimized configuration using the 72 discretization. In addition, the

optimized configuration from the 72 discretization produces more power than Horns Rev in

both wind direction resolution cases.

109
Chapter 8 Conclusions and future work

A Wind Farm Performance Evaluation Model is developed in this dissertation to

evaluate wind speed deficit and turbine power for large wind farms. The Jensen wake

model and the Ainslie wake model are applied to predict wake losses for Horns Rev wind

array and Nysted wind array. Predictions of wind speed deficits and turbine power losses

are evaluated and compared to observations and also other wind farm analyzing tools such

as WAsP, WindFarmer and NTUA etc. Results show that the LWAP model gives very good

predictions on both wind speed deficit and turbine power output on multiple wind cases. In

general, our predictions with Ainslie wake model are better than those with Jensen wake

model. Compare to other models, our evaluation results are better than engineering model

WAsP and as good as CFD based models such as WindFarmer and NTUA. In addition, our

model applies a general and straightforward multiple wake mode and uses simple

computational schemes. Thus, LWAP computational times are much less than those CFD

based model and LWAP can be readily applied to different array configurations and wind

farm layout optimization problem.

The LWAP model is also applied in WALOM with a Genetic Algorithm to analyze the

wind farm layout optimization problem. Wind turbine placement is optimized by a given site

information and wind distribution, so that the wake effects are minimized and therefore the

expected power production is maximized.

In the case studies of extension of previous approach, three different cases are

optimized and recomputed. It shows that in Case 1, 48 turbines can be placed with the

objective function value of 1.3164×10-3 and an efficiency of 98.056%. This is far better than

110
previous studies. In Case 2 and Case 3, 32 and 35 wind turbines obtain optimal objective

function values of 1.5749×10-3 and 8.3852×10-4. Both of them are better than previous

results. Results for Case 1 of different spacing limits and area sizes were obtained. These

two factors affect the optimal fitness value and, thus, the spatial configuration of wind

turbines in the array. It suggests that when designing a wind farm, consideration of site are

conditions (size and shape) turbine spatial proximity limits and wake models should be

made.

The second wind farm layout optimization approach by WALOM have been

proposed for real wind farm site such as Horns Rev and applies real site information and

wind distribution. Ainslie’s eddy viscosity wake model is employed. Wind data including the

Weibull distribution, turbulence intensities and wind turbine characteristics are used. The

effects of wind direction discretization sectors are studied to evaluate the reliability of

optimization results. However, the optimized results are highly affected by discretization of

the wind data. More improved resolution for wind discretization is, thus, recommended for

wind array optimization method. Results show that for realistic wind fields with variable

direction and speed distributions, the performance increases for the optimized array will

generally be large compared to a rectangular evenly space array. However, it is concluded

that the optimized results are highly affected by discretization of the wind data. More

improved resolution for wind discretization is, thus, recommended for wind array

optimization method.

The LWAP Model has shown a reasonable degree of agreement with the data sets

from Horns Rev and Nysted wind. Other wake models are recommended to be applied in

111
future work such as Larsen’s wake model and the Fuga wake model. It is also suggested that

multiple wakes model used in LWAP could be extended to consider wake effects that

originate further upstream.

It has been shown that factors such as wind turbine spacing limits, site area sizes and

wind data resolution can strongly affect the results of wind array performance and array

optimization calculations. Other factors such as complex terrain and wind turbine type will

also be important and should be evaluated in future work.

In addition, current optimized wind turbine placements seem to be very random and

subject to a lot of uncertainty. As a result, more optimization results are needed and more

case studies are suggested to be in the future work.

112
Bibliography

[1] Global wind report – annual market update 2010, GWEC 2011.
[2] Renewables 2011 global status report, Paris, REN21 Secretariat, 2011.
[3] Fthenakis V, Kim H C. Land use and electricity generation: A life-cycle analysis.
Renewable and Sustainable Energy Reviews 2009; 13: 1465.
[4] World Wind Energy Report 2010. World Wind Energy Association. February 2011.
Retrieved 8-August-2011.
[5] Sawin J L, Wind Power Increase in 2008 Exceeds 10-year Average Growth Rate.
Worldwatch.org. Retrieved 29 August 2010.
[6] BTM Forecasts 340-GW of Wind Energy by 2013. Renewableenergyworld.com. 27 March
2009. Retrieved 29 August 2010.
[7] BTM Consult. International Wind Energy Development World Market Update, 2009.
[8] Barthelmie RJ, Hansen K, Frandsen ST, Rathmann O, Schepers JG, Schlez W, Phillips J,
Rados K, Zervos A, Politis ES and Chaviaropoulos PK. Modelling and measuring flow and
wind turbine wakes in large wind farms offshore. Wind Energy 2009; 12: 431–444.
[9] Sanderse B, Aerodynamics of wind turbine wakes: literature review. ECN 2009. Report
ECN-E-09e016.
[10] Thomsen K, Sørensen P. Fatigue loads for wind turbines operating in wakes. Journal of
Wind Engineering and Industrial Aerodynamics 1999; 80: 121–136.
[11] Hasager CB, Rasmussen L, Peña A, Jensen LE, Réthoré PE. Wind Farm Wake: The Horns
Rev Photo Case. Energies. 2013; 6:696-716.
[12] Mechali M, Jensen L, Barthelmie R, Jensen LE, Réthoré PE. Wake effects at Horns Rev
and their influence on energy production. Wind Energy Conference and Exhibition 2006.
Athens, Greece: p. 10.
[13] Barthelmie, RJ and Jensen LE. Evaluation of wind farm efficiency and wind turbine wakes
at the Nysted offshore wind farm. Wind Energy, 2010; 13: 573-586.
[14] Patel MR. Wind and Power Solar Systems. Boca Raton: CRC Press, 1999.
[15] Ammara I, Leclerc C, Masson C. A viscous three-dimensional differential/actuator-disk
method for the aerodynamic analysis of wind farms. Journal of Solar Energy Engineering

113
2002; 124:345–56.
[16] Schlezt W and Neubert A. Special Wake Cases, presented at VindKraftNet: Wake
Meeting 16th September 2008 at Dong Energy, Copenhagen. (See also, Proceedings
DEWEK 2006 Bremen, November 2006 “New developments in precision wake
modelling”, Wolfgang Schlez, Anja Neubert, Gillian Smith).
[17] Frandsen S, Barthelmie R, Pryor S, Rathmann O, Larsen S, Højstrup J, Thøgersen M.
Analytical modelling of wind speed deficit in large offshore wind farms. Wind Energy
2006; 9: 39–53. DOI: 10.1002/we.189
[18] Frandsen S, Rathmann O, Barthelmie R, Jørgensen HE, Badger J, Hansen K, Ott S,
Rethore P, Larsen SE, and Jensen LE. The making of a 2nd generation wind farm model,
Wake Meeting 16th September 2008 at Dong Energy, Copenhagen.
[19] Larwood S. Wind Turbine Wake Measurements in the Operating Region of a Tail Vane.
NREL/CP-500-29132 2001
[20] Frandsen S, Thomsen K. Change in fatigue and extreme loading when moving wind
farms offshore. Wind Engineering 1997; 21: 197-214.
[21] Barthelmie RJ, Folkerts L, Ormel FT, Sanderhoff P, Eecen PJ, Stobbe O. Offshore wind
turbine wakes measured by SODAR. Journal of Atmospheric and Oceanic Technology
2003; 20:466-477.
[22] Christiansen MB, Hasager CB. Wake effects of large offshore wind farms identified from
satellite SAR. Remote Sensing of Environment 2005; 98: 251-268.
[23] Barthelmie R J, Larsen GC, Frandsen ST, Folkerts L, Rados K, Pryor SC, Lange B, Schepers
G. Comparison of Wake Model Simulations with Offshore Wind Turbine Wake Profiles
Measured by Sodar. Journal of Atmospheric Oceanic Technology 2006; 23: 888–901.
[24] http://www.offshorewindenergy.org
[25] Mosetti G, Poloni C, Diviacco B. Optimization of Wind Turbine Positioning in Large
Windarrays by Means of a Genetic Algorithm. Journal of Wind Engineering and Industrial
Aerodynamics 1994; 51:105-16.
[26] Jensen NO. A note of wind generator interaction, Roskilde, Denmark: Riso National
Laboratory 1993.

114
[27] Grady S. Placement of Wind Turbines Using Genetic Algorithms. Renewable Energy 2005;
30:259-70.
[28] Marmidis G, Lazarou S, Pyrgioti E. Optimal placement of wind turbines in a wind park
using Monte Carlo simulation. Renewable Energy 2008; 33:1455-1460.
[29] Mittal A. Optimization of the Layout of Large Wind Arrays Using a Genetic Algorithm.
Thesis. Case Western Reserve University, Cleveland, OH, 2010.
[30] Kusiak A, Song Z. Design of Wind Farm Layout for Maximum Wind Energy Capture.
Renewable Energy 2010; 35:685-694.
[31] Li S, Alexander J. Optimization of Wind Turbine Placement in Offshore Wind Arrays.
ASME International Design Engineering Technical Conferences 2011: Washington DC, US.
DETC2011-47935.
[32] Li S, Alexander J. Optimization of Wind Turbine Array Performance in Offshore Wind
Farms. ASME International Mechanical Engineering Congress & Exposition 2012:
Houston, TX, US. IMECE2012-87901.
[33] Wang F, Liu D, Zeng L. Study on computational grids in placement of wind turbines using
genetic algorithm. World Non-Grid-Connected Wind Power and Energy Conference 2009.
[34] Wan C, Wang J, Yang G, Zhang X. Optimal Siting of Wind Turbines Using Real-Coded
Genetic Algorithms. EWEC 2009.
[35] Huang H. Efficient hybrid distributed genetic algorithms for wind turbine positioning in
large wind arrays. IEEE International Symposium on Industrial Electronics 2009; iss. ISlE,:
2196-2201.
[36] Şişbot S, Turgut Ö, Tunç M, Çamdalı Ü. Optimal positioning of wind turbines on
Gökçeada using multi-objective genetic algorithm. Wind Energy 2010; 13: 297-306.
[37] Chen L, MacDonald E. A New Model for Wind Farm Layout Optimization with Landowner
Decisions. ASME International Design Engineering Technical Conferences 2011.
DETC2011-47772.
[38] Betz A. Introduction to the Theory of Flow Machines. (D. G. Randall, Trans.) Oxford:
Pergamon Press 1966.
[39] Gijs AM, Van Kuik. The Lanchester–Betz–Joukowsky Limit. Wind Energy 2007; 10:289–

115
291.
[40] Katic I, Hojstrup J, Jensen NO. A simple model for cluster efficiency. Proceedings of the
European Wind Energy Conference and Exhibition 1986; 407-410.
[41] Mortensen NG, Heathfield DN, Myllerup L, Landberg L, Rathmann O. Wind Atlas Analysis
and Application Program: WAsP 8 Help Facility. Risø National Laboratory, Roskilde,
Denmark.2005; 335 topics. ISBN 87-550-3457-8.
[42] Larsen GC, Madsen HA, Niels N. Mean wake deficit in the near field. European Wind
Energy Conference and exhibition 2003, Madrid, Spain.
[43] Schlichting H. Boundary layer Theory. McGraw-Hill Book Company 1968, Sixth edition.
[44] Ishihara T, Yamaguchi A, Fujino Y. Development of a New Wake Model Based on a Wind
Tunnel Experiment. Global Wind Power, 2004.
[45] Werle MJ. A New Analytical Model for Wind Turbine Wakes. FloDesign Inc., Wilbraham,
MA, 2008.
[46] Lissaman PBS. Energy effectiveness of arbitrary arrays of wind turbines. American
Institute of Aeronautics and Astronautics, 17th Aerospace Sciences Meeting, New
Orleans, 1979.
[47] Templin RJ. An estimation of the interaction of windmills in widespread arrays. National
Aeronautical Establishment. Lab. Report LTR-LA-171, Ottawa, 1974.
[48] Newman BG. The spacing of wind turbines in large arrays. Energy Conversion, 1977; 16:
169-171.
[49] Bossanyi EA, Whittle GE, Dunn PD, Lipman NH, Musgrove PJ, Maclean C. The efficiency
of wind turbine clusters. Proc. International Symposium on Wind Energy Systems,
Lyngby, 1980; 3: 401-416.
[50] Frandsen S. On the wind speed reduction in the center of large clusters of wind turbines.
Journal of Wind Engineering and Industrial Aerodynamics, 1992; 39: 251-265.
[51] Emeis S, Frandsen S, Reduction of horizontal wind speed in a boundary layer with
obstacles, Boundary- Layer Meteorology, 1993; 64: 297-305.
[52] Crespo A, Hernandez J, Frandsen S. Survey of Modelling Methods for Wind Turbine
Wakes and Wind Farms. Wind Energy; 1999: 1-24.

116
[53] Sforza PM, Stasi W, Smorto M, Sheerin P. Wind turbine generator wakes, American
Institute of Aeronautics and Astronautics, 17th Aerospace Sciences Meeting, New
Orleans, 1979.
[54] Taylor PA. On wake decay and row spacing for WECS farms. 3rd International
Symposium on Wind Energy Systems, Lyngby, 1980; 451-468.
[55] Liu MK, Yocke MA, Myers TC. Mathematical model for the analysis of wind-turbine
wakes. Journal of Energy 1983; 7: 73-78.
[56] Crasto G, Gravdahl AR. CFD wake modeling using a porous disc. European Wind Energy
Conference and Exhibition 2008, Brussels, Belgium.
[57] Ainslie J. Calculating the flowfield in the wake of wind turbines. Journal of Wind
Engineering and Industrial Aerodynamics 1988; 27: 213-224.
[58] Magnusson M. Near wake behavior of wind turbines. Journal of Wind Engineering and
Industrial Aerodynamics 1999; 80: 147-167.
[59] Rathmann O, Barthelmie RJ, Frandsen ST. Wind turbine wake model for wind farm
power production, European Wind Energy Conference, Athens, 2006.
[60] Barthelmie RJ, Pryor S, Frandsen ST, Hansen KS, Schepers JG, Rados K, Schlez W,
Neubert A, Jensen LE, Neckelmann S. Quantifying the Impact of Wind Turbine Wakes on
Power Output at Offshore Wind Farms. Journal of Atmospheric and Oceanic Technology
2010. 27: 1302-1317.
[61] "WindFarmer Website", Garrad Hassan & Partners Ltd, Accessed 2004, URL:
www.garradhassan.com/windfarmer/windfarmer.htm.
[62] Schlez W, Neubert A. New developments in large wind farm modelling. Proc. European
Wind Energy Conference, Marseilles, France, European Wind Energy Association, 2009.
[63] Schepers JG. ENDOW: Validation and improvement of ECN’s wake model. ECN-C-03-034
113 pp, 2003.
[64] Magnusson M, Rados KG, Voutsinas SG. A study of the flow down stream of a wind
turbine using measurements and simulations. Wind Eng. 1996; 20: 389–403.
[65] Politis ES, Rados K, Prospathopoulos JM, Chaviaropoulos PK, Zervos A. CFD modeling
issues of wind turbine wakes under stable atmospheric conditions. Proc. European Wind

117
Energy Conference 2009, Marseille, France, European Wind Energy Association, PO.163.
[66] Frandsen ST. Turbulence and turbulence generated structural loading in wind turbine
clusters, Risø-R-1188(EN), 2007.
[67] Dekker JWM, Pierik JTG, editors. European Wind turbine standards II.
[68] Anderson M. Simplified Solution to the Eddy- Viscosity Wake Model. Document
prepared by Renewable Energy Systems Ltd (“RES”), 2009.
[69] Crespo A, Hernandez J, Fraga E, Andreu C. Experimental validation of the UPM computer
code to calculate wind turbine wakes and comparison with other models. Journal of
Wind Engineering and Industrial Aerodynamic 1988; 27: 77–88.
[70] Schepers G, Barthelmie R, Rados K, Lange B, Schlez W. Large off-shore windfarms:
Linking wake models with atmospheric boundary layer models. Wind Engineering 2001,
25:307–316.
[71] Nielsen M, Jørgensen H, Frandsen S. Wind and wake models for IEC 61400-1 site
assessment. Wind Energy, EWEC 2009.
[72] Sempreviva AM, Larsen SE, Mortensen NG, Troen I. Response of neutral boundary layers
to changes of roughness. Boundary-Layer Meteorology 1990; 50:205-225.
DOI:10.1007/BF00120525.
[73] Kristensen L, Rathmann O, Hansen SO. Extreme winds in Denmark. Journal of Wind
Engineering and Industrial Aerodynamics, 2000. 87; 147-166.
[74] Larwood S. Wind Turbine Wake Measurements in the Operating Region of a Tail Vane.
NREL/CP-500-29132 2001.
[75] http://wis.usace.army.mil/wis.shtml
[76] Wind Statistics and the Weibull Distribution. Wind-power-program.com. Retrieved 11
January 2013.
[77] Eltra PSO-2000 Proj. NR. EG-05 3248. Wind Resources at Horns Rev, Program for
measuring wind, wave and current at Horns Rev, 2002
[78] Jensen L, Mørch C, Sørensen P, Svendsen K. Wake measurements from the Horns Rev
wind array. Elsam Engineering A/S Kraftvaerksvej 53, 7000 Fredericia.

118
[79] Davis L, Steenstrup M. Genetic search and simulated annealing. Lawrence Davis and
Morgan Kauffman, Los Altos, 1987.
[80] Goldberg, David E. Genetic Algorithms: in Search, Optimization & Machine Learning,
Addison-Wesley Publishing Company, Inc. New York, 1989.
[81] Nino R, Eecen P. ECN-DOWEC: Turbulence and wind shear: A literature study and
measurements. DOWEC- FIWI- PE-01-044/00, 2002.
[82] Barthelmie RJ, Frandsen ST, Rathmann O, Hansen KS, Politis E, Prospathopoulos J,
Schepers JG, Rados K, Cabezon D, Schlez W, Neubert A, Heath M. Flow and wakes in
large wind farms: Final report for UpWind WP8. Roskilde: Danmarks Tekniske
Universitet, Risø Nationallaboratoriet for Bæredygtig Energi. (Denmark.
Forskningscenter Risoe, 2011. Risoe-R; No. 1765(EN)).
[83] Ozturk UA, Norman BA. Heuristic methods for wind energy conversion system
positioning. Electric Power Systems Research; 70: 179-185, 2004.
[84] Christopher NE, James F, Mcgowan JG. Offshore Wind Farm Layout Optimization
(OWFLO) Project: Preliminary Results, Renewable Energy, p. 1-9.
[85] Bryony L, Jonathan C. An extended pattern search approach to wind farm layout
optimization. ASME International Design Engineering Technical Conferences &
Computers and Information in Engineering Conference IDETC/CIE 2010; p. 1-10.
[86] Christiansen MB, Hasager CB. Wake effects of large offshore wind farms identified from
satellite SAR. Remote Sensing of Environment 2005; 98: 251-268.
[87] Renkema DJ. Validation of Wind Turbine Wake Models Using Wind Farm Data and Wind
Tunnel Measurements. Master of Science Thesis. Delft University of Technology, 2007.
[88] Duckworth A, Barthelmie RJ. Investigation and Validation of Wind Turbine Wake Models.
Wind Engineering 2008; 32: 459-475.
[89] Kristensen L, Rathmann O, Hansen SO. Extreme winds in Denmark. Journal of Wind
Engineering and Industrial Aerodynamics 2000; 87:147-166.
[90] Larsen A, Larose GL, Livesey FM. Wind engineering into the 21st century. (pp. 243-250).
Rotterdam: Balkema Publishers, A.A. / Taylor & Francis The Netherlands, 1999.
[91] Frandsen ST, Barthelmie RJ, Rathmann O, Ejsing Jørgensen H, Badger J, Hansen KS, Ott S,

119
Rethore, P-E M, Larsen SE, Jensen LE. Summary Report: The Shadow effect of large wind
farms: measurements, data analysis and modelling: Risø-R-1615 (EN). (1 ed.) Risø,
Roskilde, DK: Risø National Laboratory, 2007. (Denmark. Forskningscenter Risoe. Risoe-R;
No. 1615(EN)).
[92] Milborrow DJ. The performance of arrays of wind turbines. Journal of Wind Engineering
and Industrial Aerodynamics 1980; 5:403-430.
[93] Hansen KS, Barthelmie RJ, Jensen LE, Sommer A. The impact of turbulence intensity and
atmospheric stability on power deficits due to wind turbine wakes at Horns Rev wind
farm. Wind Energy 2012; 15:183–196. doi: 10.1002/we.512.
[94] Pryor SC, Barthelmie RJ. Climate change impacts on wind energy: A review. Renewable
and Sustainable Energy Reviews 2010; 14:430-437.
[95] Rathmann O, Frandsen S, Nielsen M. Wake Decay Constant for the Infinite Wind Turbine
Array, EWEC 2010 Techn. Track “Wake”, paper ID 265.

120

You might also like