enhanced layout optimization and wind aerodynamic models ... · enhanced layout optimization and...

92
Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design by Yen Jim Kuo A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Mechanical & Industrial Engineering University of Toronto c Copyright 2016 by Yen Jim Kuo

Upload: others

Post on 20-Mar-2020

30 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Enhanced Layout Optimization and Wind Aerodynamic Models for WindFarm Design

by

Yen Jim Kuo

A thesis submitted in conformity with the requirementsfor the degree of Doctor of Philosophy

Graduate Department of Mechanical & Industrial EngineeringUniversity of Toronto

c© Copyright 2016 by Yen Jim Kuo

Page 2: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Abstract

Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design

Yen Jim Kuo

Doctor of Philosophy

Graduate Department of Mechanical & Industrial Engineering

University of Toronto

2016

The proposed project is motivated by the need to develop wake models and optimization algorithms that

can accurately capture the wake losses in an array of wind turbines and optimize the turbine placements.

In the past 4 years, we have developed capabilities to improve the layout design of wind farms located

on complex terrains, as contributions from four major tasks.

The outcome of the first task was the creation of a wake interaction model capable of describing the

effects of overlapping wakes that can be used in combination with existing mathematical optimization

tools for wind farm layout design. Such a model was derived and evaluated against existing wake

interaction methods. This wake interaction model enables a mechanistic approach to account for multiple

overlapping wakes while remaining compatible with established mathematical optimization methods.

In the second task, this wake interaction model was used in conjunction with full-scale CFD simula-

tions to design wind farm layouts. We developed an optimization algorithm that intelligently integrates

a mathematical optimization approach to design wind farm layout on complex terrains with full-scale

CFD simulations.

The two subsequent tasks were focused on developing a wake model capable of producing comparable

accuracy as full-scale CFD simulations but at a significantly lower computational cost. The third task

focused on studying the effects of turbine blade geometry and atmospheric turbulence on turbine wake

development. The findings of this step contributed to the fourth task of developing a new wake model

capable of simulating wakes on complex terrains. This model has been validated against full-scale CFD

simulations of a turbine placed on the terrain of the Gros-Morne Wind Farm in Quebec. The proposed

model allows for fast simulation of wakes, making it ideal for designing wind farm layouts on complex

terrains.

ii

Page 3: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

To Charlie Matsubara

iii

Page 4: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Acknowledgements

Firstly, I would like to express my most sincere gratitude to my supervisor Professor Cristina Amon

for her steadfast support of my Ph.D study. She worked tirelessly to provide invaluable guidance while

I pursued my academic interests. I could not have imagined having a better mentor and role model for

my Ph.D study.

Another supportive mentor and role model is Dr. David Romero. I am very fortunate to have a

mentor with such dedication and commitment to lifelong learning. He never ceases to amaze me with

the depth and breadth of his knowledge and experience. Our discussions have shaped my research and

also my life perspective in general.

Besides my immediate mentors, I would like to thank the rest of my supervisory committee: Professor

Timothy Chan and Professor David Sinton for their insightful comments and encouragement. Professor

Chan introduced me to a new world of math that was completely foreign to me, while Professor Sinton

kept me true to my academic roots. Their suggestions and thought-provoking questions incented me to

widen my research from various perspectives.

I am extremely grateful to Professor Amy Bilton and Professor Jerzy Maciej Floryan for agreeing to

be my examiners and for their thorough feedback on my thesis.

My sincere thanks also goes to Professor Joaquin Moran and Professor Deborah Tihanyi, who pro-

vided me invaluable opportunities in teaching.

Of course, I would also like to thank Helen Ntoukas and Brenda Fung, as they worked tirelessly

behind the scenes to provide professional and timely support that is second to none.

There is so much more to graduate school than just research. I have been fortunate enough have

met a wonderful group of like-minded people, Peter, Auyon, Sami, Taewoo, Kimia, Derya, and Olivier.

Although everyone will eventually take different paths in life, the friendships that were forged at UofT

will be ironclad. Furthermore, I want to thank the University of Toronto Operations Research Group

(UTORG) for providing a platform for me to connect with fellow UTORGers, particularly Anna, Carly,

Chris, Curtiss, Dexter, Heyse, Justin, Marlee, Philip, Rachel, Sarina, Shefali, Tony, and Vahid.

I thank my fellow labmates for the all the stimulating discussions, all the fun, and all the late nights.

As the ATOMS office became my professional home, the lab members also became my professional

family – Aditya, Aydin, Armin, Carlos, Danyal, David, Enrico, Fernan, Francisco, Juan, Julia, Matthew,

Patrick, Ping, Ryan, Sam, Sami, Sean, Weiguan, and Wendy.

Last but not the least, I owe my deepest gratitude and love to my family, who supported me uncon-

ditionally through this journey, my parents, my brother, Helen, Muagee, and Charlie.

iv

Page 5: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Contents

Bibliography 1

1 Introduction 1

2 Literature Review 3

2.1 Wind Farm Optimization Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2.2 Wake Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.2.1 Single Wake . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.3 Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

3 Multiple Turbine Wake Interactions 9

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

3.2 Wake Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3.2.1 Single Wake Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3.2.2 Wake Interaction Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3.3 Proposed Wake Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

3.3.1 Energy Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

3.3.2 Model Fitting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3.4 Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

3.4.1 Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

3.5 Description of Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

3.6 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

4 Layout Optimization on Complex Terrains 26

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

4.2 Previous Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

v

Page 6: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

4.2.1 Optimization Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

4.2.2 CFD Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

4.3 Proposed WFLO Optimization Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

4.3.1 MIP Optimization Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

4.3.2 Wake Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

4.3.3 Impact of the Initial Wake Approximation . . . . . . . . . . . . . . . . . . . . . . . 32

4.4 Case Study: The Carleton-sur-Mer Wind Farm . . . . . . . . . . . . . . . . . . . . . . . . 33

4.4.1 Initial Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4.4.2 Manipulating the Relaxation Parameter . . . . . . . . . . . . . . . . . . . . . . . . 36

4.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

5 Influence of Rotor Geometry and Atmospheric Turbulence 48

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

5.2 Wind Turbine Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

5.3 Model Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

5.4 Numerical and Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

5.5 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

5.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

6 Wake Model for Complex Terrains 58

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

6.1.1 Wind Turbine Wake . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

6.1.2 Actuator Disk Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

6.2 Proposed Wake Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

6.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

6.3.1 Turbulent Viscosity Model Fitting . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

6.3.2 Complex Terrains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

6.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

7 Future Work 68

Bibliography 69

vi

Page 7: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

List of Tables

3.1 Wake interaction models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

3.2 Wind turbine parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.3 Layout of 1 x 20 domain for a land strip of 2 km long. The x coordinates [m] for turbines

T1–T5 and the resulting annual energy production (AEP) [GWh] are shown. . . . . . . . 20

3.4 Layout of 1 x 100 domain for a land strip of 2 km long. The x coordinates [m] for turbines

T1–T5 and the resulting annual energy production (AEP) [GWh] are shown. . . . . . . . 21

3.5 Annual energy production [GWh] for WR6 10 x 10 on a 4 km x 4 km domain. The best

solution found for each case is indicated in boldface type. . . . . . . . . . . . . . . . . . . 22

3.6 Forfeited annual revenue for different wake interaction methods with WR6 10 x 10 on a 4

km x 4 km domain, assuming an electricity price of $0.1/kWh [1, 2]. The best solutions

found (Table 3.5) are used as reference values. . . . . . . . . . . . . . . . . . . . . . . . . . 22

3.7 Annual energy production [GWh] for WR36 20 x 20 on a 4 km x 4 km domain. The best

solution found for each case is indicated in boldface type. . . . . . . . . . . . . . . . . . . 25

3.8 Forfeited annual revenue for different wake interaction methods with WR36 20 x 20 on a

4 km x 4 km domain, assuming an electricity price of $0.1/kWh [1, 2]. The best solutions

found (Table 3.7) are used as reference values. . . . . . . . . . . . . . . . . . . . . . . . . . 25

4.1 Influence of relaxation parameter on solution quality and computational cost . . . . . . . 38

5.1 Mesh sensitivity analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

vii

Page 8: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

List of Figures

1.1 Global map of wind speeds at 80 m above the ground, courtesy of Vaisala [3]. . . . . . . . 2

2.1 Turbine wakes in Horns Rev Wind Farm in Denmark [4]. . . . . . . . . . . . . . . . . . . 3

2.2 Wind turbine wake velocity recovery. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2.3 Wake regions of a wind turbine. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

3.1 Two overlapping turbine wakes, inlet A of a streamtube is upstream in the free stream

and outlet B is in the wake overlap. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

3.2 Layout of Horns Rev wind farm, arranged in 8 rows and 10 columns. Arrows show wind

directions and the corresponding spacing, expressed in rotor diameters [D]. . . . . . . . . 14

3.3 Wind speeds experienced by a row of turbines separated by 7 diameter distances apart.

Comparison between measurements (Horns Rev) and the proposed model (with Jensen

and Frandsen) are shown. Error bars represent the standard deviation of the measurements. 15

3.4 Wind speeds experienced by a row of turbines separated by 9.4 diameter distances apart.

Comparison between measurements (Horns Rev) and the proposed model (with Jensen

and Frandsen) are shown. Error bars represent the standard deviation of the measure-

ments. Note that the experimental data exhibits a non-monotonic behavior that cannot

be captured by the single-wake models commonly used in the literature [5, 6]. . . . . . . . 16

3.5 Wind speeds experienced by a row of turbines separated by 10.4 diameter distances apart.

Comparison between measurements (Horns Rev) and the proposed model (with Jensen

and Frandsen) are shown. Error bars represent the standard deviation of the measurements. 17

3.6 Simple wind farm domain divided into 36 cells under a one-directional wind regime. The

distance between cell center is five times the rotor diameter. The wake of turbine placed

in cell j propagates downwind to affect cell i. Consequently, the placement of turbine in

cell j affects the decision whether to place a turbine in cell i or not. . . . . . . . . . . . . 18

3.7 A row of 5 turbines with a constant wind blowing from the west. . . . . . . . . . . . . . . 19

viii

Page 9: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

3.8 Layouts for the case of WR6 and 50 turbines with 10 x 10 grid and different interaction

models, LS, SKED, present work (Horns Rev), and present work (α = 1). . . . . . . . . . 23

3.9 Layouts for the case of WR36 and 40 turbines with 20 x 20 grid and different interaction

models, LS, SKED, present work (Horns Rev), and present work (α = 1), from left to right. 24

4.1 Flowchart of the optimization algorithm process . . . . . . . . . . . . . . . . . . . . . . . 29

4.2 Turbine wake created by west wind. The wake from turbine at location i propagates

downstream, affecting location j. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

4.3 Actuator disk model by El Kasmi and Masson [7]. . . . . . . . . . . . . . . . . . . . . . . 32

4.4 2.8 km x 2.8 km wind farm domain in Carleton-sur-Mer . . . . . . . . . . . . . . . . . . . 34

4.5 Wind rose for Carleton-sur-Mer. [8] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

4.6 Wind farm domain for CFD simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

4.7 Optimal layout found at the end of each iteration. The circles mark the turbines that

were relocated in that iteration. Note that after only 3 iterations, the algorithm did not

identify additional turbine locations that would lead to improvements in the optimization

objective. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4.8 Progression of the objective values (varying the relaxation factor) based on wake effects

known at each iteration. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

4.9 Optimal layout found at the end of each iteration with relaxation parameter, C, set to

0.7. The circles mark the turbines that were relocated in that iteration. Note that after

only 3 iterations, the algorithm did not identify additional turbine locations that would

lead to improvements in the optimization objective. . . . . . . . . . . . . . . . . . . . . . 40

4.10 Optimal layout found at the end of each iteration with relaxation parameter, C, set to

0.4. The circles mark the turbines that were relocated in that iteration. Note that after

only 3 iterations, the algorithm did not identify additional turbine locations that would

lead to improvements in the optimization objective. . . . . . . . . . . . . . . . . . . . . . 41

4.11 Optimal layout found at the end of each iteration with relaxation parameter, C, set to

0.2. The circles mark the turbines that were relocated in that iteration. Note that after

only 3 iterations, the algorithm did not identify additional turbine locations that would

lead to improvements in the optimization objective. . . . . . . . . . . . . . . . . . . . . . 42

4.12 Optimal layout found at the end of each iteration with relaxation parameter, C, set to

0. The circles mark the turbines that were relocated in that iteration. Note that after 8

iterations, the algorithm did not identify additional turbine locations that would lead to

improvements in the optimization objective. . . . . . . . . . . . . . . . . . . . . . . . . . . 44

ix

Page 10: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

4.13 Effects of relaxation parameter on computational cost (fraction of maximum number of

CFD evaluations). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

4.14 Effects of relaxation parameter on layout efficiency. . . . . . . . . . . . . . . . . . . . . . . 46

5.1 An actuator disk model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

5.2 Axial induction factor along a LM8.2 turbine blade. . . . . . . . . . . . . . . . . . . . . . 50

5.3 Simulation domain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

5.4 Velocity profile 1 rotor diameter downstream of the turbine with turbulence intensities of

5% and 10%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

5.5 Velocity profile 2 rotor diameters downstream of the turbine with turbulence intensities

of 5% and 10%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

5.6 Velocity profile 6 rotor diameters downstream of the turbine with turbulence intensities

of 5% and 10%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

5.7 Velocity profile 1 rotor diameter downstream of the turbine with different force profiles

at a turbulence intensity of 5%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

5.8 Velocity profile 2 rotor diameters downstream of the turbine with different force profiles

at a turbulence intensity of 5%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

5.9 Velocity profile 6 rotor diameters downstream of the turbine with different force profiles

at a turbulence intensity of 5%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

5.10 Velocity profile 1 rotor diameter downstream of the turbine with different force profiles

at a turbulence intensity of 10%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

5.11 Velocity profile 2 rotor diameters downstream of the turbine with different force profiles

at a turbulence intensity of 10%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

5.12 Velocity profile 6 rotor diameters downstream of the turbine with different force profiles

at a turbulence intensity of 10%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

6.1 Near and far wake regions of a wind turbine . . . . . . . . . . . . . . . . . . . . . . . . . . 59

6.2 Actuator disk velocity and pressure distribution . . . . . . . . . . . . . . . . . . . . . . . . 59

6.3 Actuator disk theory, the wake experiences an initial wake expansion in the near wake. . . 61

6.4 Streamwise velocity contour for both CFD and proposed model. . . . . . . . . . . . . . . . 62

6.5 Increase in turbulence viscosity inside a turbine wake. . . . . . . . . . . . . . . . . . . . . 64

6.6 Streamwise velocity profiles on a flat terrain with new turbulent viscosity fitting. . . . . . 65

6.7 CFD domain with terrain based on Gros-Morne Wind Farm. . . . . . . . . . . . . . . . . 66

6.8 Streamwise velocity at hub height. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

6.9 Streamwise velocity on a complex terrain using CFD and proposed model. . . . . . . . . . 67

x

Page 11: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 1

Introduction

Wind energy has experienced tremendous growth in the past two decades, driven by government policies

and falling costs of wind energy [9]. In 2014, the European Union set a legally binding target that by the

year 2030, at least 27% of the energy consumption must come from renewable sources [10]. According

to European Wind Energy Association, of the electricity generated by renewables, nearly half will come

from wind energy [10]. Similarly, Canada and the United States have both experienced and expect to

continue to experience the trend of double-digit growths in their respective wind energy industries in

the next decade [11, 12].

In fact, wind energy resource potential in North America is among the best in the world [13]. Ac-

cording to Lu et al. [13], the wind energy potential in the United States is more than 5 times that of its

annual energy consumption. In terms of total wind energy potential by country, Russia, Canada, United

States, Australia, and China are at the top of the list, with vast majority of that resource located inland.

By observing the global map of wind energy resource, shown in Figure 1.1, it is clear that higher wind

speeds are seen near coastal and mountainous regions.

The wind energy resource potential of a site is the single most important factor in determining the

future performance of a wind farm. Consequently, much of the work available in the literature is focused

on using scientific forecasting methods to find optimal locations for potential wind farm sites. Wind

turbines produce power by extracting energy from the air and they can be arranged into wind farms.

Commercial wind farms are typically made up of dozens of large wind turbines placed on areas that span

over several kilometers. One of the key aspects of wind farm design is the placement of wind turbines,

as aerodynamic losses, or wake losses, are one of the major sources of inefficiency in wind farms [14].

The wind farm layout optimization (WFLO) problem is concerned with the optimal placement of

turbines in a geographical area to maximize energy production and minimize costs. In order to optimize

turbine placements, aerodynamic wake models capable of describing these wake losses need to be used in

1

Page 12: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 1. Introduction 2

Figure 1.1: Global map of wind speeds at 80 m above the ground, courtesy of Vaisala [3].

conjunction with optimization approaches to evaluate turbine layouts. Previously, much effort has been

focused on developing aerodynamic models [5, 6, 15, 16, 17] and optimization approaches [18, 19, 20,

21, 22, 23] for flat and uniform terrains; however the wind energy development in the United States and

Canada has been concentrated inland [24]. Therefore, there is a need to develop wake models capable of

simulating wake effects and optimization approaches to optimize wind farm layouts on complex terrains.

The focus on this work is to develop wake models capable of describing the wake effects and applying

mathematical optimization tools to wind farm design. In Chapter 3, a model that accounts for multiple

wake interactions is presented. This model allows for a mechanistic approach to describe multiple wake

interactions that is compatible with existing mathematical optimization methods. In Chapter 4, the

wake interaction model was used in conjunction with computational fluid dynamics (CFD) simulations

to design wind farm layouts. We proposed an optimization algorithm that intelligently integrates CFD

simulation data with a mathematical optimization approach to design wind farm layouts on complex

terrains.

The following two Chapters are focused on developing a wake model with comparable accuracy with

CFD simulations with significantly lower computational cost. Chapter 5 focuses on studying the effects

of turbine blade geometry and atmospheric turbulence on turbine wake development, leading to a new

wake model capable of simulating wakes on complex terrains, described in Chapter 6.

Page 13: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 2

Literature Review

2.1 Wind Farm Optimization Problem

The main objective of a WFLO problem is to maximize energy production while minimizing costs. The

power production of a wind farm is dependent on incoming wind speeds, which are themselves dependent

on terrain topography, atmospheric conditions, and upstream turbine wakes. In particular, production

loss due to the wake interference of upstream turbines, called wake losses (Figure 2.1), can reduce the

annual energy of a wind farm by as much as 10% to 20% [14].

Figure 2.1: Turbine wakes in Horns Rev Wind Farm in Denmark [4].

3

Page 14: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 2. Literature Review 4

Wind turbines extract kinetic energy from the wind via interactions between the turbine blades and

the wind. The aerodynamic forces produced by the wind turns the rotor to generate electricity. The air

behind the turbine is slowed down with its turbulence intensity increased [16]. This region of decelerated

air is called the wake [16]. The wake expands as it travels downstream, mixing with surrounding air

and increasing its velocity back to undisturbed conditions after some distance. This distance is crucial

as the performance of turbines downstream is dependent on the incoming wind conditions. If turbines

are too close to each other, the wind cannot recover to its upstream state [25], causing losses in energy

generation for turbines downstream [14, 17, 26]. Therefore, a better understanding of turbine wake

recovery process is crucial for optimal wind farm design and planning [25].

The main purpose of this review is to present the relevant state-of-the-art tools for designing wind

farms on complex terrains. Most of the development related to optimization and wake modelling have

focused on flat and uniform terrains. However, some of these important advances can be adapted to

design wind farms on complex terrains. This review examines key optimization approaches and wake

models available in the literature.

This chapter is divided into two main sections, existing wake models and optimization approaches.

The wake model section is divided into single wake and multiple wake interaction models. The single

wake model section discusses a variety of models, ranging from low-fidelity analytical models to full-scale

computational fluid dynamics models. While single wake models are well-developed, models that can

describe the interactions of multiple overlapping wakes is limited in the literature. Overall, we found that

the phenomena of wake interactions is not well understood, suggesting research gaps and opportunities

to be pursued.

From the design optimization perspective, the field of optimizing turbine placement is relatively

mature. In the literature, a wide selection of optimization approaches have been applied to solve WFLO

problems. These established optimization approaches usually rely on analytical wake models for solution

evaluation. However, analytical wake models do not account for effects on complex terrains; this makes

their use limited. Based on these observations, we developed a set of tools to bridge the gap between

wake models and optimization approaches to determine wind turbine placement on complex terrains.

2.2 Wake Models

2.2.1 Single Wake

A turbine wake expands and propagates downstream, mixes with the mean flow, then its velocity recovers

over large distances, shown in Figure 2.2. Wake models are used to quantify this wake recovery and

propagation process.

Page 15: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 2. Literature Review 5

Figure 2.2: Wind turbine wake velocity recovery.

Wind turbine wakes can be divided into two regions: near wake and far wake regions [27, 28]. The

near wake region begins from the turbine to a specific location and then is followed by the far wake

region as shown in Figure 2.3. The length of the near wake region varies, ranging from 1 to 4 diameters

downstream of the rotor [29, 30]. Previous studies have shown that the velocity and turbulence profiles

in the near wake region are strongly influenced by the rotor geometry [31, 32]. In the far wake, the

influence of rotor geometry becomes less important while terrain and atmospheric effects become more

dominant. In addition, the velocity profile becomes approximately Gaussian and the pressure gradient

due to the turbine becomes negligible [27]. The far wake region is the region of interest for wind farm

layout design, as turbines are typically placed far apart such that they are in the far wake regions of

upstream turbines [33].

Figure 2.3: Wake regions of a wind turbine.

Far wake models are divided into two categories, kinematic and field models [28]. Kinematic wake

models such as Jensen [5], Frandsen [6], Ishihara [34], and Larsen [35] are analytical models that can be

solved very quickly, making them ideal for optimization. Field models such as Ainslie [15] use a reduced

form of the Reynolds-averaged Navier-Stokes equations with added turbulence model.

Page 16: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 2. Literature Review 6

In addition, with advances in computing power, numerical wake models have been introduced and

implemented in commercial software packages. For instance, Ainslie [15] first developed the eddy viscosity

wake model, which was implemented in the WindFarmer software [36]. Then, similar implementations

of the model was done in OpenWind [37] and The Farm Layout Program (FLaP) [38]. The Energy

Research Centre of the Netherlands introduced a wake model based on the parabolic Navier-Stokes

equations using k − ε turbulence model [39]. These models specifically simulate wakes on flat terrains

and cannot be used to model wake effects on complex terrains.

Full-scale computational fluid dynamics (CFD) models have been applied to study wind turbine

wakes [14, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50]. Full-scale CFD simulations involve solving the full

Navier-Stokes equations. One of the main challenges in modelling turbine wakes is how the turbine rotor

should be represented in CFD simulations. In order to reduce the computational cost, the actuator

disk model [51] is used to describe the complex rotor geometries. The results of this approach has

shown to produce good agreement with full rotor modelling approach [52]. Jafari et al. [53] proposed a

novel approach in which actuator disk theory is used to describe the wake expansion and velocity and

turbulence profiles are prescribed at the end of the near wake region.

The main challenge in developing wake models for complex terrains is reducing the computational

cost. During the design optimization process, wake models are used to evaluate the wake effects of

different layouts. This process is iterative and requires a large number of solution evaluations, thus low

cost wake model is necessary to evaluate large pools of candidate solutions. In commercial software,

wake models for complex terrains rely on variations of wake superposition approaches [54]. Recently, a

virtual particle model [55] has also been developed to simulate wakes on complex terrains. This model

utilizes the concept of solving for the velocity deficit instead of velocity directly by assuming velocity

deficit is analogous to concentration of particles. However, due to the assumptions, this model requires

parameter tuning that cannot be known a priori nor be generalized.

Multiple Wake Interactions

The interaction of multiple superimposed wakes is not fully understood, as it involves complex turbulence

phenomena [56]. A number of descriptions exist in the literature to determine the wind speed due to the

presence of multiple turbine wakes. In particular, four descriptions available in the literature [16, 57, 58].

These descriptions are recursive functions, which is dependent on conditions upstream which is not

known a priori. However, these functions are not mechanistic in nature, lacking physical basis, which

makes improvement through experimental results difficult. Furthermore, these descriptions are non-

linear, leading to non-linear objective functions which makes it challenging to use with well-established

mathematical programming approaches [18, 19, 59]. In Chapter 3, a physics-based wake interaction

Page 17: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 2. Literature Review 7

model that leads to linear objective functions is presented.

2.3 Optimization

The main objective of a wind farm layout optimization (WFLO) problem is to maximize energy pro-

duction while minimizing costs such as installation and maintenance costs. Determining the optimal

layout of a set of wind turbines in a given area is a complex problem. Wind farm layout problems

can be modelled in two ways, namely (1) continuous and (2) discrete. In discrete models [20, 21, 60],

the turbines can only be placed in a countable set of pre-determined locations inside the wind farm,

while in continuous models [61, 62, 63, 64, 65, 66, 67], turbines can be placed anywhere in the farm,

considering their coordinates as continuous variables. Metaheuristic algorithms such as evolutionary al-

gorithms [60, 64, 65, 66, 67], particle swam optimization [63, 68], and extended pattern search [62], have

been the primary tools to solve continuous models. Although powerful in tackling non-linear problems,

metaheuristics cannot guarantee global optimality.

A discrete model can be solved by using mathematical programming approaches, which are promising

in solving WFLO problems [19, 59, 69, 70, 71, 72]. Donovan [59, 73] introduced a mixed-integer pro-

gramming (MIP) model for solving the WFLO problem. Although commercial MIP solvers are widely

available and are capable of proving optimality of solutions, they are limited to solving convex and

linear or quadratic problems. Both Donovan and Fagerfjall [70] attempted to address this problem by

simplifying their wake model at the expense of losing accuracy. To address this issue, Archer et al.

[74] improved the simplified wake model by introducing a wind interference coefficient, while Turner et

al. [19] suggested more accurate linear and quadratic wake models that can be solved by MIP solvers.

Furthermore, Zhang et al. [69] proposed a constraint programming model that incorporates the full

non-linearity of the problem. Although these approaches allows to proof of optimality, they rely on

the relatively coarse discretization of wind farm domains, because the solution time required to solve

increase exponentially with finer discretizations.

On the other hand, continuous models are solved using evolutionary metaheuristic algorithms [60,

62, 63, 64, 65, 66, 66, 67, 68, 75, 76] and nonlinear optimization methods [71, 77]. A significant portion

of the literature has focused on improving the model by including realistic constraints and features. For

instance, Yamani Douzi Sorhabi et al. [78] studied the impact of land use constraints on the impact

of noise and energy in a multi-objective optimization. Furthermore, Serrano-Gonzalez et al. [79, 80]

included infrastructure costs and wind data uncertainty in their optimization model. While Saavedra-

Moreno et al. [81] improved the model by considering multiple wind distributions for different locations

in the wind farm domain. Although the WFLO problem refers specifically to the design optimization

Page 18: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 2. Literature Review 8

of turbine placements [82, 83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95, 96, 97, 98, 99, 100, 101, 102]

sometimes the number of turbines, turbine type [103], and turbine height [95, 104, 105] are considered

in the optimization problem [106].

Most of the publications in the literature consider the terrains to be uniform and flat, ignoring

the effects of terrain elevation. Terrain topography strongly influences the local energy potential of a

farm as well as the wake propagation and recovery process. The work by Saavedra-Moreno et al. [81]

considered the effects of terrain topography on local wind speed but did not account for them in the

wake propagation and recovery process. The major hurdle is that the wake propagation and recovery

process on complex terrains is difficult to model accurately and cost-effectively for optimization. Feng

and Shen [54] assumed that the wind turbine wake propagates along a complex terrain at hub height.

Then, Song et al. [55] introduced a wake model based on particle simulations and integrating it with

various optimization algorithms [95, 105, 107, 108, 109] to design wind farms on complex terrains. Given

the reciprocal relationship between optimization approaches and wake models, it is important to consider

this dependency during their respective development, and this is the main objective of this thesis.

Page 19: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3

Multiple Turbine Wake Interactions

3.1 Introduction

There are a number of publications on discrete modelling of wind turbine placement in wind farms

[20, 21, 110, 111]. An example is the work by Mosetti et al. [20], where the wind farm is divided into 10

by 10 square cells and each turbine is placed in the centre of each cell. The cell sizes are chosen to enforce

distance constraints between turbines, e.g., turbines cannot be placed closer than five turbine rotor

diameters apart. Although layout solutions to discrete models may be of lower spatial resolution than of

continuous models, discrete models can be solved using powerful mathematical programming approaches

[19, 59, 69], which can guarantee optimality of the solutions for linear and quadratic functions and

constraints. In a well-designed discrete model, knowing the optimality of solutions can save tremendous

amount of time in the optimization process.

A mixed integer programming problem (MIP) consists of an objective function and a mix of integer

and continuous constraints. The layout optimization problem can be modelled in this mathematical

programming approach by discretizing the wind farm domain into possible turbine placement locations,

with binary decision variables denoting if a turbine is placed at a specific location or not. These problems

can be solved using algorithms such as branch and bound [112]. Fagerfjall [70], Donovan et al. [59], Zhang

et al. [69], and Turner et al. [19] have studied the application of branch and bound methods in WFLO

problems. Applying mathematical programming methods to solve the layout problem is promising due

to the optimality of the solutions can be known, as opposed to metaheuristics that provide no guarantee

of convergence. Furthermore, solver efficiency can be improved through alternative problem formulations

and problem-specific branching strategies [69]. The objective of this chapter is to introduce a novel wake

interaction model that leads to linear MIP formulations, thus guaranteeing the optimality of solutions

to WFLO problems.

9

Page 20: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 10

3.2 Wake Modeling

3.2.1 Single Wake Model

The Jensen model [5] is one of the most widely used wake models. It assumes a linearly expanding wake

and uniform velocity profile inside the wake. As a result of momentum conservation, the decelerated

wind behind the rotor recovers to free stream speed after travelling a certain distance downstream of

the turbine [5]. The velocity downstream from the rotor is given by

u(x) = u0

[1− 1−

√1− CT

(1 + 2k xD )2

], (3.1)

where CT is the thrust coefficient of the turbine, D is the turbine rotor diameter, u0 is the wind speed

in the free stream, and k is the Wake Decay Constant, which is generally taken to be 0.075 for onshore

farms and 0.04 to 0.05 for offshore farms.

The power production of each turbine i is based on the incoming wind speed that it experiences,

Pi =1

2ρAu3i (ηgenCP ), (3.2)

where A as the rotor area, ρ is the density of the air, ηgen is the generator efficiency, and CP is the rotor

power coefficient. The annual energy production (AEP) of a wind farm is defined as the integration of

power production (kW) over time (hr),

AEP = 8766

N∑i=1

∑d∈L

pdPi,d. (3.3)

where pd is the probability of wind state d, defined as a (speed, direction) pair, L is the set of wind

states with non-zero probability for the specific wind farm site, N is the total number of turbines, and

8766 is the effective number of hours in a year.

3.2.2 Wake Interaction Models

The interaction of multiple superimposed turbine wakes is not fully understood, as it involves complex

turbulence phenomena [56]. A number of descriptions exist in the literature to determine the wind speed

due to the presence of multiple turbine wakes upstream. In particular, four descriptions available in the

literature [16], listed in Table 3.1, will be introduced in this section. In these equations, ui is the wind

speed at turbine i, uij is the wind speed at turbine i due to (the wake of) turbine j and the summations

and the products are taken over the n turbines upstream of turbine i [16, 57, 58].

The Geometric Superposition (GS) assumes the ratio of the wind speed at a location relative to the

Page 21: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 11

free stream speed is a product of velocity ratios caused by upstream turbines. The Linear Superposition

of Velocity Deficits (LSVD) considers that the velocity deficit at a given turbine is equal to the sum of

the velocity deficits caused by all turbines upstream from it. The Sum of Energy Deficits (SED) assumes

the kinetic energy deficit in the wakes is additive. The Sum of Squares (SS) sums up the squares of the

velocity deficits of the upstream wakes.

Table 3.1: Wake interaction models

Name Formula

Geometric Superposition (GS) uiU∞

=∏nj=1

uijuj

Linear Superposition of Velocity Deficits (LSVD)(

1− uiU∞

)=∑nj=1

(1− uij

uj

)Sum of Energy Deficits (SED)

(U2∞ − u2i

)=∑nj=1

(u2j − u2ij

)Sum of Squares (SS)

(1− ui

u∞

)2=∑nj=1

(1− uij

uj

)2Two main issues exist that hinder the use of these wake interaction models. Firstly, with the ex-

ception of SED, the physical basis of these descriptions is unclear, which makes improvement through

experimental data difficult. Secondly, using all of the mentioned wake interaction models in deterministic

optimization methods remains a challenge, for several reasons. If the objective function of the WFLO

problem is to maximize power or energy production, all of the above wake interaction models (except

SED) would lead to non-linear objective functions and linear approximations will be required to improve

solvability. In addition, all of the models mentioned are recursive functions. Specifically, the term uj , i.e.

the incoming wind speed that a turbine j experiences, is dependent on the conditions upstream, which

are unknown a priori, thus precluding the use of well-established mathematical programming methods

for its solution [19, 59, 69].

In the literature, comparisons of the different wake interaction models with experimental measure-

ments have demonstrated that the sum of squares model, despite lacking physical meaning, is the most

accurate [16]. However, as mentioned previously, it is difficult to implement sum of squares into a MIP

formulation. In this work, a wake interaction model based on the principle of energy balance is presented

as a physics-based, linear alternative to the sum of squares model, leading to linear objective functions.

3.3 Proposed Wake Model

3.3.1 Energy Balance

In the proposed wake interaction model, energy balance is used to describe the wind speed recovery in

the far wake, where the flow is fully developed [25]. The speed recovery in the wake is due to mixing

with the surrounding air, causing changes in kinetic energy, this change is represented by a mixing head,

Page 22: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 12

Figure 3.1: Two overlapping turbine wakes, inlet A of a streamtube is upstream in the free stream andoutlet B is in the wake overlap.

h. Without loss of generality, consider a simple case of two wakes overlapping, shown in Figure 3.1.

Energy analysis is done along a streamtube from A to B, ignoring the presence of the bottom turbine.

The mixing head hAB becomes

hAB(1) = PA1 + αA1u202g− PB1 − αB1

u2B1

2g, (3.4)

where αA1 and αB1 are kinetic energy correction factors, accounting effects of nonuniform speed profiles

in the streamtube. The other terms P , u, and g represent local pressure, wind speed, and gravity,

respectively. In a flat terrain, the pressure in the wake is assumed to have recovered to mean flow

pressure, thus PA1 = PB1. However, if pressure changes need to be captured, the pressure terms could

be kept easily in the analysis. The mixing head hAB can be simplified to

hAB(1) = αA1u202g− αB1

u2B1

2g, (3.5)

Similarly, the same analysis can be done again from A to B, ignoring the top turbine, leading to

hAB(2) = αA2u202g− αB2

u2B2

2g. (3.6)

Page 23: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 13

When the wakes of two turbines overlap, we assume in this work that the mixing gains in the combined

wake is equal to the energy loss in the free stream, thus the analysis from A to B becomes

α∞u202g

= αBu2B2g

+∑

h, (3.7)

or

u2B =α∞αB

u20 +

n∑j

(αBjαB

u2Bj −αAjαB

u20), (3.8)

where n is the total number of overlapping wakes at point B. The assumption that uA ≈ u0 may

not hold if an upstream turbine is too close to downstream turbines, e.g., for wake interactions when

turbines are densely placed together. Specifically, our computer experiments showed that the proposed

model outperformed benchmarks for wind farms in which the inter-turbine spacing was larger than seven

turbine rotor diameters (7D), with relative performance degrading (but still comparable) in the 5–7D

range, presumably because the experimental data used to calibrate the proposed model corresponds to

a wind farm in which the closest turbines are a distance of 7D apart. In any case, however, we note that

our assumption of uA ≈ u0 is not expected to be a limitation because an inter-turbine spacing constraint

is typically enforced during wind farm design and typical wind turbine densities (turbines per square

kilometre) found in existing wind farms such as Horns Rev and Nysted [113].

The kinetic energy correction factors can be determined experimentally, and we considered them in

this work as model fitting parameters to be estimated based on available data from real wind farms [14].

Consequently, experimental data can be used to improve the accuracy of the model. Based on turbulent

pipe flows, these values of these coefficients should be close to 1 [114]. To simplify the analysis, the

ratios α∞αB

,αBjαB

, andαAjαB

are denoted as αr,1, αr,2, and αr,3, respectively, assuming these coefficients are

constant in the far wake region for all j. Equation (3.8) becomes

u2i = αr,1u20 +

n∑j

(αr,2u2ij − αr,3u20). (3.9)

If no experimental or detailed CFD (computational fluid dynamics) data is available, the model can

be used as a surrogate model, or metamodel [115, 116, 117] for the SS model, in which the model is a

linear approximation of SS. The coefficients can be obtained with synthetic data generated from direct

evaluation of the SS model, an approach that will be explored in future work. In addition, the coefficients

could be set to 1 based on considerations that are typically valid in turbulent pipe flows. In this work,

we will compare the performance of wind farm layouts that are determined with the proposed model

both when the coefficients are regressed to data from the Horns Rev wind farm [118] and also when they

Page 24: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 14

are considered as constants set to 1.

3.3.2 Model Fitting

The coefficients of the proposed model are determined using experimental data from Horns Rev wind

farm in Denmark. Horns Rev wind farm is made up of 80 2 MW turbines in a structured layout of

8 rows and 10 columns, as shown in Figure 3.2. Publicly available data from wake measurement at

wind directions indicated in Figure 3.2, are used for parameter fitting. The wind speeds along a row of

turbines are presented as a function of distance between upwind and downwind turbines. Figures 3.3,

3.4, and 3.5 show the wind speeds in a row of 5 turbines separated by constant distances of 7, 9.4, and

10.4 diameters at 6 m/s, respectively. The error bars are the standard deviations of the measurements

of the available rows [118].

Figure 3.2: Layout of Horns Rev wind farm, arranged in 8 rows and 10 columns. Arrows show winddirections and the corresponding spacing, expressed in rotor diameters [D].

Based on the Horns Rev data shown in Figures 3.3–3.5, the model coefficients are found to be

αr,1 = 0.936, αr,2 = 0.9375, and αr,3 = 0.8885, and the fitted wake interaction model, Eq. (3.9), is

also shown in the figures. These coefficients minimize the root mean square error when comparing with

experimental data. The proposed model performs within the error bars for 7 and 10.4 diameter distances,

while the wind speeds predicted by the model do not fall within the error bars for 9.4 diameter distances.

A closer examination of the experimental data showed that the wind speed recovered faster for the 9.4D

case than for the 10.4D case, a behavior that does not correspond to what the single-wake Jensen model

would predict. Consequently, this non-monotonic and faster-than-expected recovery for the 9.4D case

Page 25: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 15

is not an artifact of the proposed wake interaction model, but it is instead attributable to the Jensen’s

model inability to reproduce the observed wake recovery.

It is important to demonstrate that the proposed wake interaction model is suitable for other wake

models other than the Jensen model. As a comparison, the Frandsen wake model [6] is used with the

proposed wake interaction model, and the results of model fitting are also shown in Figures 3.3–3.5. Note

that the proposed wake interaction model fits the experimental data within its error bars, indicating

that it can be used to model wake interactions regardless of the approach used to model the single-wake

speed recovery.

Figure 3.3: Wind speeds experienced by a row of turbines separated by 7 diameter distances apart.Comparison between measurements (Horns Rev) and the proposed model (with Jensen and Frandsen)are shown. Error bars represent the standard deviation of the measurements.

3.4 Optimization

3.4.1 Model

The wind farm layout problem can be formulated as a mixed integer programming (MIP) problem. The

wind farm domain is partitioned into a set of cells, with the restriction that each cell can have at most

one turbine placed in its geometric centre. A simple wind farm under a simple one-directional wind

regime shown in Figure 3.6 is discretized such that inter-turbine distance constraint of 5 rotor diameters

Page 26: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 16

Figure 3.4: Wind speeds experienced by a row of turbines separated by 9.4 diameter distances apart.Comparison between measurements (Horns Rev) and the proposed model (with Jensen and Frandsen)are shown. Error bars represent the standard deviation of the measurements. Note that the experimentaldata exhibits a non-monotonic behavior that cannot be captured by the single-wake models commonlyused in the literature [5, 6].

is enforced. The proposed MIP formulation, similar to that proposed by Donovan [59, 119] and Zhang et

al. [69], will be in the form of kinetic energy deficit (equivalent to the summation portion of Eq. (3.9)),

meaning that the objective of the formulation would be to minimize the kinetic energy loss (mixing

head). Let xi be a binary decision variable, indicating whether a turbine is placed (xi = 1) or not

(xi = 0) in cell i, i = 1, .., n. Let pd be the probability of wind state d, where a wind state is defined

as a (speed, direction) pair, and let L be the the total number of wind states with non-zero probability.

Then, the optimization formulation can be written as

Page 27: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 17

Figure 3.5: Wind speeds experienced by a row of turbines separated by 10.4 diameter distances apart.Comparison between measurements (Horns Rev) and the proposed model (with Jensen and Frandsen)are shown. Error bars represent the standard deviation of the measurements.

min

n∑i=1

∑d∈L

xi∑j∈J

(hij)xjpd (3.10a)

s.t

n∑i=1

xi = k (3.10b)

Qx ≤ 1.5 (3.10c)

xi ∈ {0, 1} ∀i = 1, ..., n (3.10d)

(3.10e)

where the first constraint describes the total number of wind turbines k to be placed in the wind farm

domain, which is held constant. In Eq. (3.10c), Q is a binary matrix that is calculated prior to the

optimization as an aid to enforce the inter-turbine distance constraints. For example, if cells i and j are

closer than 5 rotor diameters apart, a row would be generated in the matrix Q with 1’s in the i-th and

j-th columns and 0’s everywhere else.

Also, in Eq. (3.10a), the hij term denotes the kinetic energy deficit (mixing head) at cell i caused

by a turbine at cell j. This is found by using the summation term in Eq. (3.9), hij = −αr,2u2ij +αr,3u20.

Page 28: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 18

As previously described, the coefficients αr,2 and αr,3 are used to fit the wake interaction model to

experimental measurements. The Jensen wake model is used to determine the wind speed for each

individual wake.

Figure 3.6: Simple wind farm domain divided into 36 cells under a one-directional wind regime. Thedistance between cell center is five times the rotor diameter. The wake of turbine placed in cell jpropagates downwind to affect cell i. Consequently, the placement of turbine in cell j affects the decisionwhether to place a turbine in cell i or not.

3.5 Description of Tests

Three sets of tests were conducted to evaluate different wake interaction models used for MIP in the

literature. The MIP formulation was chosen since the optimality of the solutions can be determined,

allowing for a fair and accurate comparison between the underlying wake interaction models. The first

set of tests is aimed at illustrating the importance of wake interaction models for layout optimization.

The second set of tests aims to study how the current proposed model performs against existing wake

interaction models, namely the Linear Superposition (LS) method used by Zhang [69] and the Sum of

Kinetic Energy Deficits (SKED) approximation by Turner et al. [19]. The last set of tests is intended to

assess the quality of the near-optimal solutions produced from each model if optimal solutions cannot

be found in the time allotted for the optimization run.

The first set of tests determines the effects of wake interaction models in WFLO problems by studying

the placement of 5 turbines on a horizontal land strip that is 2 km long and 100 m wide, under a constant

wind of 6 m/s blowing from west to east. Due to the land’s narrow width, the turbines can only be placed

downstream of each other, as shown in Figure 3.7. To test the effect of the discretization resolution on

the resulting wind farm layouts, two discretization resolutions are tested, dividing the 2 km x 0.1 km

wind farm domain into 20 cells and 100 cells. The turbine parameters for all tests are shown in Table

Page 29: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 19

3.2.

Figure 3.7: A row of 5 turbines with a constant wind blowing from the west.

The second set of tests is conducted to evaluate the performances of different wake interaction models

(LS and SKED) and against the proposed physics-based, linear model with coefficients determined from

Horns Rev data. In addition, to assess the applicability of the proposed model when no measurement

data is available, we also studied the performance of the proposed model with all coefficients set to 1,

i.e., αr,1 = αr,2 = αr,3 = 1. These test cases involved a simple wind regime and a varying number of

turbines, using a 4 km x 4 km square wind farm domain, discretized into 10 x 10 cells, each cell is a

square of size 400 m x 400 m. The wind regime includes six equally probable wind directions at 60 ◦

increments starting from the north, with a constant wind speed of 6 m/s. This wind regime is denoted

as WR6. The turbine details are the same as in the previous test set, as shown in Table 3.2. Note that

the size of these problems is sufficiently small such that they can be solved to optimality in a reasonable

amount of time, so that any observed differences in performances can only be attributed to the wake

interaction models, rather than lack of convergence to their respective optimal solutions [120].

The final test is performed to evaluate the ability for the models to produce good near-optimal

solutions. One of the advantages that mathematical programming optimization methods provide is the

known optimality of solutions. However, in large complex problems, it may not be possible to solve them

to optimality in sufficient time. Thus, the ease to solve a particular model becomes very important. In

this test, turbines are placed in a 4 km x 4 km square domain, divided into 20 x 20 cells with WR36

wind regime, where the wind blows at a constant wind speed of 6 m/s from 36 directions.

The optimization model was implemented using MATLAB and Gurobi 5.6 on a Dell Poweredge T420

Tower Server, 8 Intel Xeon processor E5-2400, and 164 GB RAM. The Gurobi linearization function to

convert quadratic integer problems to mixed-integer linear problems was used. Then, the problems

are solved through a branch-and-bound process. The total simulation time limit was set to 384 hr of

CPU time. In order to ensure a fair comparison with other results reported in the literature, all the

optimization results are re-evaluated with the sum of squares model, regardless of the model that was

used to obtain the results.

Page 30: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 20

Table 3.2: Wind turbine parameters

Parameter Value

Rotor Diameter 80 mThrust Coefficient 0.805Power Curve 1.68u3 kWWake Decay Constant 0.05

Table 3.3: Layout of 1 x 20 domain for a land strip of 2 km long. The x coordinates [m] for turbinesT1–T5 and the resulting annual energy production (AEP) [GWh] are shown.

Model T1 T2 T3 T4 T5 AEP

Present Work 50 450 1050 1550 1950 8.17

LS 50 450 950 1450 1950 8.15

SKED 50 450 950 1550 1950 8.12

3.6 Results and Discussion

The first test case illustrates the importance of wake interaction models. Based on intuition about

problem behavior, it is clear that first and last turbines would be placed in the first and last cells,

leaving the remaining 3 turbines to be placed in the rest of the domain. This simple case also served as

a validation check for the problem formulation.

The layouts found highlight the influence and importance of wake interaction models. Table 3.3 shows

the optimal turbine positions for 1 x 20 domain discretization. The influence of the wake interaction

models on turbine positions is apparent. For example, the proposed model places the third turbine 100

m further downstream than SKED and LS, while the fourth turbine is placed in the same location as

SKED. In this simple case, the differences in AEP found using different models are noticeable, with the

proposed model outperformed the others.

The wind farm domain is further discretized into 1 x 100 cells to test the influence of discretization

resolution. The results are shown in Table 3.4. The differences in turbine layout positions for the three

wake interaction models have narrowed but are still noticeable. In terms of AEP, all three layouts have

similar performance as the problem is too restrictive, leaving turbines little room to move to improve

performance. For this simple case, it is clear that the solutions are grid dependent.

Typically, smaller problems require lower computation cost, thus optimal solutions can be found very

quickly with MIP formulations. For example, the optimal layouts for the first set of tests (1 x 20 domain)

were found in less than 0.05 seconds (wall clock time) for each case. However, as the number of cells

increases, the difficulty of the problem increases exponentially [121]. Thus, it may not be possible to

solve the problem to optimality in a timely fashion for finer grids. For example, in the 1 x 100 domain,

optimal solutions were reached in less than 5 seconds, a 100-fold increase even though the problem size

increased only 5-fold.

Page 31: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 21

Table 3.4: Layout of 1 x 100 domain for a land strip of 2 km long. The x coordinates [m] for turbinesT1–T5 and the resulting annual energy production (AEP) [GWh] are shown.

Model T1 T2 T3 T4 T5 AEP

Present Work 10 410 1010 1590 1990 8.34

LS 10 410 990 1530 1990 8.33

SKED 10 450 990 1530 1990 8.34

In the second set of tests, the performance of proposed model, LS model, and SKED model are

evaluated under WR6 wind regime, for the problem of placing 40 to 70 turbines in the 4 km x 4 km

wind farm domain; note that Turner et al. [19] showed that cases with these numbers of turbines are

particularly challenging to solve. Table 3.5 shows the resulting AEP for this case, comparing the LS, the

SKED, and the proposed wake interaction model with two different sets of model coefficients, namely

(a) coefficients fitted to data and (b) coefficients set to 1. The Gurobi solver was able to solve all the

problem instances to optimality within a wall-clock time of 200 seconds.

In Table 3.5, the best solutions found are indicated in boldface type, note that the proposed model

outperforms both the LS and SKED wake interaction models. Moreover, even when the proposed model

coefficients are set to 1, i.e., assuming that there is no experimental data available to calibrate the model,

the proposed model outperforms the others.

In order to relate the AEP values shown in Table 3.5 with economic gains/losses, Table 3.6 shows

the annual revenues that would be forfeited if the proposed model was not used. The layout found

using proposed model with Horns Rev coefficients produces additional revenue of $26,000–458,000 USD

compared with other models, assuming a wind energy price of approximately $0.10 USD/kWh [1, 2].

Even when the proposed model is used with nominal coefficients, not fitted to any experimental data,

larger AEP values, and corresponding financial gains, can still be realized. It is important to note that

the AEP values were calculated using a wind speed of only 6 m/s, which is considered to be a low

wind speed when compared with the rated wind speed of large wind turbines. Of course, the absolute

differences in AEP will be higher at higher speeds. For example, the amount of revenue forfeited by

using the SKED model (instead of our proposed model with nominal coefficients of α ≡ 1) in the case

with 70 turbines would increase from $26K at a 6 m/s wind speed to $62K at 8 m/s and $120K at 10 m/s.

The optimized layouts found with different wake interaction models for 50 turbines under the WR6

wind regime are shown in Figure 3.8. The LS layout (Figure 3.8a) is the worst performing, with the two

top rows and bottom occupied while layouts from SKED (Figure 3.8b) and proposed model (Figures

3.8c and 3.8d) occupied the top and bottom rows and spreading the turbines out in the remaining

domain. Overall, the results for the second test case show that the proposed model produces better

results than existing wake interaction models and that choosing an appropriate wake interaction model

Page 32: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 22

Table 3.5: Annual energy production [GWh] for WR6 10 x 10 on a 4 km x 4 km domain. The bestsolution found for each case is indicated in boldface type.

Number ofTurbines

LS SKEDPresent Work(Horns Rev)

Present Work(α ≡ 1)

40 114.97 114.97 114.97 114.97

50 125.82 128.75 129.51 129.03

60 137.13 141.71 141.71 141.71

70 148.45 149.25 149.51 149.25

Table 3.6: Forfeited annual revenue for different wake interaction methods with WR6 10 x 10 on a 4 kmx 4 km domain, assuming an electricity price of $0.1/kWh [1, 2]. The best solutions found (Table 3.5)are used as reference values.

Number ofTurbines

LS SKEDPresent Work(Horns Rev)

Present Work(α ≡ 1)

40 - - - -

50 $370K $75K - $48K

60 $458K - - -

70 $106K $26K - $26K

is very important for layout optimization.

The third set of tests was aimed to determine how the models would perform when they’re not solved

to optimality, within the allowed simulation time limit of 384 CPU hours. Under the WR36 wind regime,

20 to 70 turbines were placed in the wind farm domain. The layouts found for 40 turbines are shown in

Figure 3.9. In these layouts, the turbines arranged themselves along the perimeter of the domain and

spread out in the interior. The AEP values and the annual revenue forfeited due to using different layouts

produced from LS, SKED, and proposed model are shown in Table 3.7 and Table 3.8, respectively. In

the WR36 wind regime, as the number of turbines increase, the annual revenue forfeited increases due

to the increased wake interactions. The present model with coefficients of 1’s outperformed that of LS

and SKED when the number of turbines range from 20 to 50, but not for 60 and 70 turbines. On the

other hand, the present model with coefficients from Horns Rev outperformed all other models. The

solutions found in this case are not globally optimal but the results demonstrated that under resource

constraints, the proposed model can produce better solutions compared with existing models.

3.7 Conclusions

In the present work, a new physics-based wake interaction model that leads to linear MIP formulations

was introduced. The proposed linear wake interaction model was compared with existing wake interaction

Page 33: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 23

(a) Optimal layout found using LS model (b) Optimal layout found using SKED model

(c) Optimal layout found using present work(Horns Rev)

(d) Optimal layout found using present work (α =1)

Figure 3.8: Layouts for the case of WR6 and 50 turbines with 10 x 10 grid and different interactionmodels, LS, SKED, present work (Horns Rev), and present work (α = 1).

Page 34: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 24

(a) Optimal layout found using LS model (b) Optimal layout found using SKED model

(c) Optimal layout found using present work(Horns Rev)

(d) Optimal layout found using present work (α =1)

Figure 3.9: Layouts for the case of WR36 and 40 turbines with 20 x 20 grid and different interactionmodels, LS, SKED, present work (Horns Rev), and present work (α = 1), from left to right.

Page 35: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 3. Multiple Turbine Wake Interactions 25

Table 3.7: Annual energy production [GWh] for WR36 20 x 20 on a 4 km x 4 km domain. The bestsolution found for each case is indicated in boldface type.

Number ofTurbines

LS SKEDPresent Work(Horns Rev)

Present Work(α ≡ 1)

20 59.61 59.76 59.92 59.80

30 86.05 85.86 86.15 85.94

40 109.96 109.73 110.63 110.13

50 131.57 131.36 132.66 131.97

60 149.75 150.67 151.46 150.23

70 164.66 165.14 165.50 165.12

Table 3.8: Forfeited annual revenue for different wake interaction methods with WR36 20 x 20 on a 4km x 4 km domain, assuming an electricity price of $0.1/kWh [1, 2]. The best solutions found (Table3.7) are used as reference values.

Number ofTurbines

LS SKEDPresent Work(Horns Rev)

Present Work(α ≡ 1)

20 $31K $16K - $12K

30 $10K $29K - $21K

40 $67K $90K - $50K

50 $110K $131K - $69K

60 $170K $79K - $122K

70 $84K $36K - $38K

models in the literature that are suitable for MIP formulations. While the interaction of multiple wakes

is not fully understood, the physics-based model can be fitted with experimental data or can be used as

a MIP-friendly surrogate model for non-linear wake interaction models, e.g., sum of squares.

Three test cases were conducted to evaluate the performance of the proposed model. The first major

finding of this study is that under the wind regime tested, better optimal layouts were found with the

proposed model compared to optimal layouts found with SKED and LS models. This result illustrated

the importance of selecting appropriate wake interaction models for WFLO problems, as they affect

energy production and consequently, wind farm economics. The second major finding is that even when

optimal solutions cannot be obtained due to resource constraints, the proposed model still outperformed

that of SKED and LS models. This has strong implications when globally optimal solutions cannot be

easily found, and near-optimal solutions are used as inputs for local search procedures to further improve

layout solutions.

The current research was not specifically designed to evaluate factors related to the capabilities of

the optimization solver. Considerable more work will need to be done to determine the performance of

these models for large complex problems and the MIP solver’s ability to find good solutions efficiently.

Page 36: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4

Layout Optimization on Complex

Terrains

4.1 Introduction

Most studies on wind farm layout optimization have focused on optimizing layouts on flat and uniform

topography [18, 19, 20, 21, 22, 23, 102]. However, wind speeds over complex terrains are very different

than they are over flat terrains, since complex flow structures can form as wind flows over various land

features. Consequently, energy production is strongly influenced by local topography. Furthermore, the

lack of analytical, closed-form mathematical models for wakes over complex terrains makes it difficult to

evaluate and optimize wind farm layouts. As a result, Feng and Shen [54] modified an adapted Jensen

wake model to estimate the wake effects of a wind farm on a two-dimensional Gaussian hill. Taking a

different approach, the virtual particle model developed by Song et al. [55] modeled the turbine wake

as concentration of non-reactive particles undergoing a convection-diffusion process in a relatively low-

cost model that describes the wake more accurately than a modified flat terrain wake model. Despite

these efforts, reducing the computational cost of wake evaluations while maintaining accuracy during

the optimization process remains a challenge. Hence, subsequent work [107, 108, 109] has focused on

better integration of wake modeling and optimization algorithms.

Computational fluid dynamics (CFD) models (e.g. actuator disk and actuator line) have been devel-

oped to simulate complex wake phenomena and their interactions with terrains [40, 46, 47, 48, 49, 50, 122].

However, these simulations are expensive and must be used sparingly during the optimization process.

Deterministic optimization approaches such as mixed-integer programming (MIP) [18, 19, 69, 123,

124] have been shown to be promising in solving WFLO problems. These models can provide global

26

Page 37: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 27

solutions and optimality bounds for relatively small problems. In the MIP model, the wind farm is

divided into discrete number of turbine locations and the wake interactions are calculated in advance for

algorithms such as branch and bound [19, 59, 69, 70, 112], to be applied to solve the WFLO problem.

The objective of this chapter is to introduce an algorithm capable of integrating CFD simulation data

into the optimization process, to intelligently design wind farm layouts located on complex terrains. In

the proposed algorithm, CFD simulation data is used as input for MIP to improve the accuracy of

the wake effects. Conversely, MIP provides information on the promising turbine locations where CFD

simulations should be conducted. This two-way coupling between MIP and CFD reduces the number of

CFD simulations significantly, and in turn the computational cost. This algorithm is applied on a wind

farm domain found in Carleton-sur-Mer, Quebec, Canada. Results show that the algorithm is capable

of optimizing layouts of wind farms on complex terrains by integrating CFD simulation data into the

optimization process.

4.2 Previous Work

4.2.1 Optimization Models

A number of approaches to tackle the WFLO problem have been developed in the literature. The WFLO

problem can be modeled by two approaches, discrete and continuous. In discrete models [20, 21, 60], the

wind farm domain is divided into a number of possible turbine locations, while for continuous models

[61, 62, 63, 64, 65, 66, 67], the turbine location is represented by two-dimensional continuous coordinates.

Continuous models are typically solved using evolutionary metaheuristic algorithms [65, 66, 75, 81, 125,

126, 127, 128, 129] and nonlinear optimization methods [71, 77]. A discrete model can be solved by using

mathematical programming approaches, which have the significant advantage of providing optimality

bounds [19, 22, 59, 69, 70].

4.2.2 CFD Models

Computational fluid dynamics models have been applied to simulate wind turbine wakes, using Reynolds-

averaged Navier-Stokes (RANS) [40, 46] and Large Eddy Simulation (LES) [26, 47, 130, 131, 132] tur-

bulence models to simulate the turbulent wake phenomena. In addition to turbulence modeling, there

are two main approaches to model rotor geometry: actuator disk/line and direct blade modeling. In

an actuator disk [40, 46, 48, 133, 134, 135] or actuator line [136, 137, 138] approach, the turbine is

modeled by imposing aerodynamic forces through a disk representing the rotor or lines representing the

turbine blades, respectively. In a direct blade modeling approach [52, 130, 139], the turbine geometries

Page 38: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 28

are inserted into the computational domain, allowing a more accurate representation of the aerodynamic

effects than the actuator disk/line approach at the expense of higher computational cost. The actuator

disk approach is less computationally expensive and less accurate. Despite the introduction of these

models in turbine wake modeling, it remains difficult to apply these models in optimization algorithms

to solve the WFLO problem due to the computational expense of CFD models.

4.3 Proposed WFLO Optimization Algorithm

While optimization and wake modeling have been applied individually to WFLO, there is a significant

challenge in combining them. An optimization algorithm typically must evaluate a very large number

of solutions and partial solutions. However, a single CFD simulation is so computationally expensive

that very few can be conducted in a reasonable run-time. In our approach, the optimization model

is first used with less accurate, less expensive data to identify promising turbine locations. The wake

effects of turbines placed at those locations are updated using CFD simulations. The CFD data is then

used iteratively by the optimization model to identify newly promising locations. Figure 4.1 shows a

schematic of our approach.

The principal idea behind the proposed algorithm is that on a complex terrain, the wind energy

potential of a location is influenced by the local terrain topography, thus different turbine locations will

have different “turbine placement potentials”. The proposed algorithm utilizes a MIP model to search

through promising locations through a combination of estimated wake effects and CFD simulation data.

Looking at the flowchart of the proposed algorithm in Figure 4.1, firstly, a flow field over the complex

terrain without turbines is generated using CFD. The initial wake effects can be calculated by superim-

posing a flat terrain wake onto the complex terrain as described in Section 4.3.2. This initial problem

is then solved to determine where the turbines should be placed. However, due to inaccuracies in the

initial wake estimate, placing turbines at these locations may not produce the optimal layout. Hence

CFD simulations are conducted at these locations to improve the accuracy of the initial estimated wake

effects. This process is repeated until no new improving turbines locations are found. In other words,

the wake effects described in the optimization model becomes more accurate with each iteration. Hence,

the optimal solution of the current iteration is more accurate than those found in previous iterations.

If the problem cannot be solved to optimality due to run-time limits, then it becomes necessary to

compare the near-optimal solutions from previous iterations. Conceivably, other optimization methods

such as metaheuristics are also compatible with this algorithm; however, without proof of optimality,

the termination criteria for the optimization problem would need to be defined appropriately.

Page 39: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 29

Figure 4.1: Flowchart of the optimization algorithm process

4.3.1 MIP Optimization Model

A number of mixed-integer programming formulations have been developed to tackle the WFLO problem

[19, 69, 123, 124]. A MIP model consists of an objective function, a set of constraints, and a mix of

integer and continuous variables. To describe the WFLO problem, the wind farm is discretized into

possible turbine locations with corresponding binary decision variables denoting if a turbine is located

at each location or not. The formulation used in this work, similar to that of the work of Kuo et al.

[123, 124], has an objective function of maximizing the sum of the kinetic energy experienced by each

turbine, as follows. Let the wind farm domain be divided into a total of N cells, let K be the number of

turbines to be placed (considered a constant in the formulation), and let xi be a binary variable denoting

whether a turbine is placed in the i-th cell. The optimization problem is

max

N∑i=1

∑s∈S

psxi

[u20,s,i −

∑j∈J

(u20,s,j − u2s,ij)xj]

(4.1a)

s.t

N∑i=1

xi = K (4.1b)

dijxi + djixj ≤ 1 ∀i, j (4.1c)

xi ∈ {0, 1} ∀i = 1, ..., N (4.1d)

Page 40: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 30

where the binary terms dij and dji indicate the violation of the distance constraint between i-th and

j-th cells, which need to be calculated in advance. Namely, dij = dji = 1 if the distance constraint

is violated when turbines are placed both in the i-th and j-th locations, and dij = dji = 0 otherwise.

In Eq.(4.1a), ps is the probability of wind state s, and S is the total number of wind states, where a

wind state is defined as a (wind speed, wind direction) pair. Most importantly, u20,s,j −u2s,ij denotes the

kinetic energy deficit at cell j caused by a turbine at cell i, which is dependent on the wind state, s.

Figure 4.2 shows a wake from turbine located in cell i, propagating downstream to affect cell j.

Figure 4.2: Turbine wake created by west wind. The wake from turbine at location i propagatesdownstream, affecting location j.

In this formulation, all the single wake effects caused by a turbine must be calculated in advance

for all possible locations. That is, when a turbine is placed in cell i, its single wake effects on all

remaining cells must be known for all possible turbine locations and wind states. Hence, the number of

potential turbine locations (i.e., the number of cells) multiplied by the number of wind states determines

the number of wake calculations required (i.e., N |S|) to define the MIP formulation. In the proposed

algorithm, the promising turbine locations are identified from the optimal MIP layout solutions using

less accurate data and CFD simulations are only conducted at these locations. In this way, we seek to

achieve the same wake accuracy as running N |S| CFD simulations with a fraction of the computational

cost.

When multiple turbines wakes are present, their combined effect on wind speed recovery is ap-

proximated by using an energy balance approach by Kuo et al. [124]. This form is suitable for MIP

formulation due to its linearity and sound physical basis. Energy balance is done along a streamtube

from the free stream mixing into the wake, assuming the wake losses are additive for overlapping wakes.

The MIP model can be solved using mathematical programming approaches to compute the optimal

turbine layout for each set of inputs.

Page 41: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 31

4.3.2 Wake Modeling

In order to identify a promising turbine placement to evaluate with a CFD simulation, we must first solve

the MIP model with approximate wake effects. These wake effects are calculated using an approximate

wake model by superimposing CFD simulation data of a flat terrain wake onto the complex terrain, using

Eq.(4.2) and Eq.(4.3). The assumptions made here are that the wake propagates downstream along the

terrain surface at hub height and that the wake will experience a speed-up factor due to terrain effects,

i.e.

uct,s,j = Ss,juft,s,j , (4.2)

uwct,s,ij = Ss,juwft,s,ij , (4.3)

where uct,s,j and uft,s,j are the free stream wind speeds on complex and flat terrains in cell j in wind

state s, and uwct,s,ij and uwft,s,ij describe the wind speeds in the wake on complex and flat terrains in cell

j due to a turbine in cell i, respectively. In other words, Ss,j , the speed-up factor due to terrain effects

experienced in cell j (in comparison with flat terrain flow field) in wind state s, is calculated without

the presence of turbines, and then used to “carry” the wakes downstream, similar to the implementation

used by Feng and Shen [54] and in several commercial software packages [54]. In this work, whenever

CFD simulation data is available, the speed-up factor Ss,j is corrected using simulation results. It should

be noted that while superimposing wakes onto terrains is not an accurate representation of the actual

wake effects, this work also addresses the effects of accuracy of initial wake approximation on solution

quality and computational cost (see following section).

When promising turbine locations are available, CFD simulations are conducted to simulate wake

effects of turbines at those locations. The actuator disk model and the extended k− ε turbulence model

by El Kasmi and Masson [7] are used in this study. Specifically, an actuator disk is inserted into the

computational domain and the turbulent dissipation zones are prescribed upstream and downstream of

the disk, shown in Figure 4.3. Appropriate boundary conditions (e.g. inlet, outlet, surface roughness)

must be prescribed to accurately simulate the atmospheric boundary layer.

To summarize how MIP and CFD are combined, the proposed algorithm is as follows:

(1) Generate flow field over the complex terrain without turbines using CFD.

(2) Construct the initial wake effects using the approximated method described in wake modelling sec-

tion.

(3) Solve the optimization problem to identify the most promising locations.

(4) Run single turbine CFD simulations at locations found in the previous step.

Page 42: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 32

(5) Update the wake effects from CFD results (us,ij term) in optimization (Eq.(4.1a)).

(6) Repeat steps (3–5) until the solution converges.

Figure 4.3: Actuator disk model by El Kasmi and Masson [7].

4.3.3 Impact of the Initial Wake Approximation

In this algorithm, the final layout is dependent on the initial wake approximation. The assumption that

wakes propagate in a straight line at the hub height may not hold for complex terrains, thus resulting in

a vast overestimate of the velocity deficit in certain cells and an underestimate in others. If the velocity

deficit is overestimated in some cells in the initial approximation, those cells might never be considered

in future layout solutions. Thus a relaxation parameter, C, is introduced to reduce the velocity deficit in

the initial wake approximation. Specifically, the velocity deficit is multiplied by the relaxation parameter,

C, to force an underestimate of the velocity deficit and mitigate the effect of poor approximations of

wake behavior on complex terrains.

When the wake effects are underestimated, more turbine locations or cells will be explored so more

CFD simulations are required. Hence, the relaxation parameter C controls how aggressively the op-

timization space is explored, balancing the need for better accuracy in wake modeling with the total

computational cost of the optimization. Specifically, the us,ij term from Eq.(4.1a) is re-written as

U0,s,j − CDs,ij , where Ds,ij is the velocity deficit at cell j caused by turbine at cell i in wind state s.

The U0,s,j −CDs,ij term is only used when CFD data is not available (these cells are defined as set N2).

If CFD data is available (defined as set N1), then the simulation data is used directly for us,ij and the

Page 43: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 33

relaxation parameter is not used. The new MIP formulation is written as,

max∑i∈N1

∑s∈S

psxi

[U20,s,i −

∑j∈J

(U20,s,j − u2s,ij)xj

](4.4a)

+∑i∈N2

∑s∈S

psxi

[U20,s,i −

∑j∈J

(2U0,s,j − CDs,ij)CDs,ijxj

](4.4b)

s.t

N∑i=1

xi = K (4.4c)

dijxi + djixj ≤ 1 ∀i, j (4.4d)

xi ∈ {0, 1} ∀i = 1, ..., N. (4.4e)

4.4 Case Study: The Carleton-sur-Mer Wind Farm

The proposed algorithm is tested on a 2.8 km x 2.8 km wind farm domain in Carleton-sur-Mer, Quebec,

Canada. The topography was extracted from Google MapsTM (https://goo.gl/maps/XTpxd), with a

roughness length assumed to be 0.1 m. The terrain elevation in meters above sea level is shown in Figure

4.4. The optimization domain is discretized into a uniform grid of 20 x 20 cells, separated at cell center

by a distance of 140 m. A wind farm layout of 20 turbines is optimized for this terrain. These turbines

are assumed to have a constant thrust coefficient of 0.8, hub height of 80 m, and rotor diameter of 80

m. These turbine parameters were obtained from the Carleton Wind Farm. The proximity constraint

between turbines is set as 5 rotor diameters apart.

For this wind farm domain, information regarding the wind speed and directions are available from

the Canadian Wind Energy Atlas [8]. A power law velocity profile is used to describe the wind speed at

varying heights

u(z) = 6(z − 139

50

)0.16, (4.5)

where z is the height above sea level. This velocity profile is used to define inlet boundary conditions for

CFD simulations. The wind rose used for this domain is shown in Figure 4.5, noting that the dominant

wind direction is from the west. The turbulent kinetic energy and dissipation rate at the inlet are

prescribed as k = (u∗)2√Cµ

and ε(z) = (u∗)3

κ(z−139) , where Cµ = 0.033 and κ = 0.4. With the assumptions for

ground roughness and the height (1000 m) of the boundary layer, the friction velocity u∗ = 0.4m/s. The

velocity and turbulence quantities are fixed at the top boundary, as other types of boundary conditions

such as symmetry or slip wall could cause undesirable streamwise gradients [48, 140]. In case the wind

Page 44: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 34

Figure 4.4: 2.8 km x 2.8 km wind farm domain in Carleton-sur-Mer

is not aligned with the x-direction, the velocity inlet takes the form of ux(z) = 6(z−139

50

)0.16cos(θ) and

uy(z) = 6(z−139

50

)0.16sin(θ), where θ is the wind direction relative to the x-axis [141]. The ground is

taken as wall boundary and the outlet face is considered as pressured outlet boundary.

4.4.1 Initial Results

To summarize the WFLO problem, 20 turbines are placed in a domain (Figure 4.4) that is discretized

into uniformly sized 20 x 20 cells. Based on the wind rose, Figure 4.5, there are 12 wind directions with

a power law wind velocity profile as given in Eq.(4.5). The proximity constraint between turbines was

set to be 5 diameters distance apart. In the initial test, the relaxation parameter has been set to C = 1.

The MIP model can be solved under 30 seconds using Gurobi 5.6, so that the bulk of the computa-

tional expense is dedicated to CFD simulations. For each cell, a CFD simulation needs to be conducted

for every wind direction, or in this case, 12 CFD simulations per cell. With 400 possible locations, and

12 wind directions, the maximum number of single turbine CFD simulations is 400 x 12 = 4800.

Each CFD simulation is performed for a domain of 2.8 km x 2.8 km in length and width, with a

height up to 1000 m above sea level, shown in Figure 4.6a. Initially, the CFD simulations are conducted

Page 45: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 35

Figure 4.5: Wind rose for Carleton-sur-Mer. [8]

without the presence of turbines for all 12 wind directions, with the domain discretized into 1.2 million

cells in the domain. When a turbine is placed in the domain, the number of cells is increased to 1.6

million cells to better capture the wake effects downstream of the turbine, shown in Figure 4.6b.

(a) CFD domain with one turbine (b) Mesh of the CFD domain.

Figure 4.6: Wind farm domain for CFD simulations

In the first iteration, the flow field in the absence of turbines is obtained from CFD simulations.

The turbine wake from flat terrain is modified using Eq.(4.3) to approximate the wake effects without

conducting any CFD wake simulations. The layout found in this first iteration is shown in Figure 4.7a.

In the second iteration, the wakes for wind turbines placed at these 20 locations are simulated using

CFD and the initial wake effects are updated. The new layout that was found is shown in Figure 4.7b.

In this new layout, three turbines are relocated compared to the first iteration. The turbine wakes from

Page 46: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 36

these three locations (indicated by circles) are simulated and updated. In the third and final iteration,

the layout found in Figure 4.7c is identical to that of the second layout, indicating that the algorithm

has converged.

(a) First iteration (b) Second iteration

(c) Third iteration

Figure 4.7: Optimal layout found at the end of each iteration. The circles mark the turbines that wererelocated in that iteration. Note that after only 3 iterations, the algorithm did not identify additionalturbine locations that would lead to improvements in the optimization objective.

4.4.2 Manipulating the Relaxation Parameter

A parametric study on the relaxation parameter, C, was conducted, considering the values C = {1, 0.7, 0.4, 0.2, 0},

to study the effects of the initial wake approximation on the solution quality and computational cost.

In a WFLO problem for complex terrains, the solution upper bound in terms of energy production is

one where the turbines are placed at locations where the wind speeds are the highest, ignoring the wake

effects. This upper bound for this test case is found to be 2177.48. Normalizing all the objective values

Page 47: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 37

found in this study with this upper bound provides a relative comparison of the solutions found using

different values of C. This normalized value is defined as the layout efficiency. The influence of the

values of C on the progression on the objective values (based on wake effects known at each iteration)

is shown in Figure 4.8.

The solutions found for different values of C are shown in Figures 4.9–4.12. The influence of C on the

number of iterations, number of CFD simulations, final objective value, layout efficiency, and run-time

is shown in Table 4.1. It can be seen that as C decreases in value, better layouts are produced. It is

notable that for the cases where C ≥ 0.2, only three iterations and a small fraction of the total number

of CFD simulations are needed for convergence. For the case of C = 0, eight iterations are required for

convergence and more CFD simulations are needed (compared with higher values of C) as the algorithm

searched through 52 turbine locations in the domain. In other words, when the wake deficits are not

accounted for, the algorithm will “blindly” search through the most promising cells in terms of wind

resource until the optimal solution is found. This behavior can be seen in Figure 4.12, where large

number of turbines are relocated to neighbouring locations from one iteration to the next, until all the

promising cells are exhausted. While computational cost is not a major concern when the size of the

problem is relatively small, and can be solved to optimality relatively quickly, this can be a significant

downside when the problem increases in size, e.g. larger number of possible turbine locations and more

complex wind regimes. For the test cases where C ≥ 0.2, the total run-time is approximately 300 hours

on a Dell Precision T1700 PC, but the run-time more than doubled when C = 0, demonstrating the

importance of the relaxation parameter in controlling the computational cost. It is important to note

that the solution found in the C = 0 case is the globally optimal solution. That is, if CFD simulation

data is available for all cell locations (N |S| = 4800 CFD simulations, approximately 5300 hours), the

optimal solution would be identical to that of C = 0, unless the presence of turbine wakes can locally

improve the energy potentials of some locations.

For all the different values of C tested, the final layout solutions are within 2% of the upper bound.

The difference in performance between the best (C = 0) and worst (C = 1) solutions is less than 1.5%,

demonstrating the algorithm’s capability to find good solutions even with poor initial estimation of

wake effects. Figures 4.13 and 4.14 show the effects of the relaxation parameter on computational cost

and layout efficiency. In terms of solution quality, underestimating the wake deficit (e.g. C = 0.2) is

desirable as the path of wake propagation is difficult to predict prior to CFD simulations. When higher

values of C are used, velocity deficits experienced by downstream turbines may be overestimated in some

cells. The consequence is that certain promising locations may be ignored during the search. However,

when wake deficits are underestimated with lower values of C, the computational cost increases. In this

particular problem, a low C value of 0.2 did not dramatically increase the computational cost relative to

Page 48: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 38

Table 4.1: Influence of relaxation parameter on solution quality and computational cost

C# of

Iterations# of CFD

Evaluations

FinalObjective

Value

LayoutEfficiency

(%)

Run-time(hr)

1 3 23 x 12 = 276 2133.25 97.97 303.63

0.7 3 22 x 12 = 264 2146.66 98.58 290.43

0.4 3 21 x 12 = 252 2150.83 98.78 277.23

0.2 3 24 x 12 = 288 2165.16 99.43 316.83

0 8 52 x 12 = 624 2165.80 99.46 686.47

larger values, but did improve solution quality significantly. Note that this algorithm is not a globally

seeking algorithm, hence the final solution is dependent on the initial layout. Based on the finding, the

relaxation factor has the effect of forcing the algorithm to converge into locally optimal solutions.

Choosing the “right” C to produce good layout will depend on the terrain topography. If the terrain

is too rugged and the flow experiences rapid changes where the streamlines can deviate significantly

from the terrain profile, a smaller C would be ideal in finding good layouts. As the local changes in

the topography is less pronounced, a larger value of C would be preferred. An intuitive and adaptive

scheme of varying values of C for every iteration can be developed, borrowing the idea from simulated

annealing [142], e.g. starting with initial low C and adjusts as the algorithm progresses. In addition,

better prediction of the initial wake effect is needed for improving solution quality and computational

cost. These two areas of improvement will be the focus in future studies.

Page 49: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 39

Figure 4.8: Progression of the objective values (varying the relaxation factor) based on wake effectsknown at each iteration.

Page 50: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 40

(a) First iteration (b) Second iteration

(c) Third iteration

Figure 4.9: Optimal layout found at the end of each iteration with relaxation parameter, C, set to 0.7.The circles mark the turbines that were relocated in that iteration. Note that after only 3 iterations,the algorithm did not identify additional turbine locations that would lead to improvements in theoptimization objective.

Page 51: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 41

(a) First iteration (b) Second iteration

(c) Third iteration

Figure 4.10: Optimal layout found at the end of each iteration with relaxation parameter, C, set to 0.4.The circles mark the turbines that were relocated in that iteration. Note that after only 3 iterations,the algorithm did not identify additional turbine locations that would lead to improvements in theoptimization objective.

Page 52: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 42

(a) First iteration (b) Second iteration

(c) Third iteration

Figure 4.11: Optimal layout found at the end of each iteration with relaxation parameter, C, set to 0.2.The circles mark the turbines that were relocated in that iteration. Note that after only 3 iterations,the algorithm did not identify additional turbine locations that would lead to improvements in theoptimization objective.

Page 53: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 43

(a) First iteration (b) Second iteration

(c) Third iteration (d) Fourth iteration

(e) Fifth iteration (f) Sixth iteration

Page 54: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 44

(g) Seventh iteration (h) Eighth iteration

Figure 4.12: Optimal layout found at the end of each iteration with relaxation parameter, C, set to0. The circles mark the turbines that were relocated in that iteration. Note that after 8 iterations,the algorithm did not identify additional turbine locations that would lead to improvements in theoptimization objective.

4.5 Conclusion

In this work, an algorithm that optimizes wind farm layouts on complex terrains was introduced. This

algorithm combines CFD simulations with mathematical programming methods for layout optimization.

To the best of the authors’ knowledge, this is the first WFLO study that makes use of mathematical

programming methods with CFD wake simulations. The proposed iterative approach identifies promising

turbine locations to minimize the number of CFD simulations required in optimization while finding good

layouts, even when the optimization relies on inaccurate wake models during the first iterations. The

proposed approach starts with an approximate wake model that superimposes a flat terrain wake model

on the topography, and this model is adaptively refined based on CFD simulations that are conducted

only at promising turbine locations. This chapter presents a better and more efficient optimization

of wind turbine layouts on complex terrain, because of better modeling accuracy and the theoretical

convergence bounds of MIP models.

In order to study the effects of initial wake approximation on solution quality and computational

cost, we introduced a relaxation parameter to control how the optimization space is explored. It was

found that regardless of the parameter value, the difference in performance for best and worst layouts

found is less than 1.5%, indicating that the algorithm is capable of finding good layouts even under poor

initial wake approximations. Finding a suitable value for the relaxation parameter will depend on the

balance between computational cost and solution quality as low values of the relaxation parameter may

Page 55: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 45

Figure 4.13: Effects of relaxation parameter on computational cost (fraction of maximum number ofCFD evaluations).

Page 56: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 46

Figure 4.14: Effects of relaxation parameter on layout efficiency.

Page 57: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 4. Layout Optimization on Complex Terrains 47

improve solution quality at the expense of computational cost while the the reverse may hold true for

high values.

Further work in developing the proposed novel approach for WFLO combining CFD simulations of

wake behavior with mathematical programming is needed to study the scalability of the algorithm to

larger problem instances, i.e., to wind farms with more potential turbine locations. The implications

of this work are that CFD can be a valuable tool in WFLO problems and that good potential turbine

locations can be identified in advance to significantly reduce the number of expensive simulations.

Page 58: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 5

Influence of Rotor Geometry and

Atmospheric Turbulence

5.1 Introduction

One of the main challenges in modelling turbine wakes is how the turbine rotor should be represented in

CFD simulations. To reduce the computational cost, the actuator disk approach [51] is used to model

the complex rotor geometries. The results from this modelling approach have been shown to produce

good agreement with the full rotor modelling approach [52].

When the details of turbine rotor geometry or its aerodynamic forces are not available, it is common

to model a turbine rotor as a disk or lines exerting a uniform force on the flow [133, 143, 144, 145]. This

approach assumes that the variable of interest is independent of the rotor geometry far downstream of

the turbine. This assumption was also made by Jafari et al. [53] and Ainslie [146] in their respective

modelling approaches. In these numerical models, the far wake is the region of interest, and the near

wake region is modelled analytically. These modelling approaches rely on two underlying assumptions:

1) the wake recovery process in the far wake region, where the atmospheric conditions are dominant,

can be determined without knowing the blade profile, and 2) the location where the far wake region

begins is relatively insensitive to blade geometry and atmospheric conditions. However, the validity of

these assumptions has not been studied in the literature. Thus, understanding the influence of turbine

geometry and atmospheric conditions on the wake recovery process is the primary objective of this study.

48

Page 59: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 5. Influence of Rotor Geometry and Atmospheric Turbulence 49

5.2 Wind Turbine Model

In an actuator disk model, the turbine is modelled as a porous disk with an infinite number of blades

that extract momentum from the flow [51, 147]. A visual representation of the model is shown in Figure

5.1. The actuator disk is a momentum sink that exerts aerodynamic forces on the incoming flow. This

thrust force (T ) can be calculated as:

T =

∫A

(p1 − p2)dA, (5.1)

where p1 and p2 represent the pressure acting on the upstream and downstream faces of the disk,

respectively, and A is the rotor swept area. This thrust force can be normalized, and the thrust coefficient

(CT ) is introduced:

CT =T

0.5ρAu20= 4a(1− a), (5.2)

where a = 1 − uwakeu0

is the axial induction factor, u0 and uwake are the wind speed in the free stream

and the wake, respectively. As previously mentioned, the aerodynamic forces that turbine blades exert

on the flow are not uniform along the radial direction. However, in the absence of detailed turbine rotor

information, these forces are assumed to be uniform.

Figure 5.1: An actuator disk model.

In this work, an actuator disk model is used in conjunction with the extended k−ε turbulence model

developed by El Kasmi et al. [41]. This k − ε model is a modification of the standard k − ε model

in which an additional term is added to more appropriately model the turbulence energy transfer rate

from large-scale to small-scale turbulence. The constants used by this model, corresponding to neutral

atmospheric conditions, are: σk = 1.0, σε = 1.3, Cε1 = 1.176, Cε2 = 1.92, Cε4 = 0.37, and Cµ = 0.033.

Page 60: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 5. Influence of Rotor Geometry and Atmospheric Turbulence 50

5.3 Model Validation

The wake model by El Kasmi et al. [41] has been validated in the literature, producing good agreement

with experimental data. The focus of this section is to validate the accuracy of the aerodynamic forces

prescribed on the actuator disk with that of wind tunnel experimental data from a test turbine. The

turbine used in this study has three LM8.2 blades with a rotor diameter of 19.06 m. The axial induction

factor, shown in Figure 5.2 [148], is used to calculate the forces acting on the incoming flow.

Figure 5.2: Axial induction factor along a LM8.2 turbine blade.

Through the use of Eq. (5.2), aerodynamic forces can be prescribed on the actuator disk, and then

the mechanical power can be calculated as:

P =

∫A

(p1 − p2)udiskdA, (5.3)

where udisk is the wind speed at the disk.

A mesh sensitivity analysis was performed to ensure mesh independence. Three different mesh sizes

were used in the validation process and their corresponding mechanical powers (calculated using Eq.

(5.3)) are shown in Table 5.1. For all three mesh sizes tested, they yielded similar mechanical power

when compared with experimental data. The error from Mesh 1 is approximately 9% with respect to

the experimental data and is the mesh used in this study.

Page 61: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 5. Influence of Rotor Geometry and Atmospheric Turbulence 51

Table 5.1: Mesh sensitivity analysis

Mesh 1 Mesh 2 Mesh 3 Measurement

Number of Elements 1.2M 4.0M 8.1M –

Power (kW) 26.34 26.42 26.47 28.97

5.4 Numerical and Experimental Setup

This main focus on this work is to study the effects of turbine blade geometry and atmospheric turbulence

on wind turbine wake development. To study the influence of blade geometry, the wake development of

an actuator disk with an aerodynamic force profile of an LM8.2 blade and that of a uniform force profile

are compared. In other words, to examine the effect of rotor geometry on wake development, two sets of

forces corresponding to an LM8.2 blade and uniform profiles that extract identical mechanical power are

prescribed for comparison. Then, the effects of atmospheric turbulence are studied under two different

conditions, turbulent intensities of 5% and 10%.

The CFD simulation domain is 24 rotor diameters in the streamwise direction and 10 rotor diameters

in both the spanwise and vertical directions. The turbine is placed 4 rotor diameters downstream of the

inlet at a hub height of 20 m. The geometry of the simulation domain is shown in Figure 5.3.

Figure 5.3: Simulation domain

A log law velocity profile is prescribed at the inlet:

U(y) =u∗

κln

(y

y0

), (5.4)

where κ = 0.4, y is the height above ground in the vertical direction, y0 is the surface roughness, and

u∗ is the friction velocity. The turbulent kinetic energy and dissipation rate at the inlet are prescribed

as k = (u∗)2√Cµ

and ε(y) = (u∗)3

κy , where Cµ = 0.033, and the friction velocity u∗ is assumed to be 0.16

Page 62: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 5. Influence of Rotor Geometry and Atmospheric Turbulence 52

m/s [141]. The roughness length y0 is prescribed as 4.036× 10−6 m and 8.984× 10−3 m, corresponding

to turbulence intensities of 5% and 10%, respectively [141]. The velocity and turbulence quantities are

fixed at the top boundary, and the bottom boundary is taken as a wall boundary, while the outlet is

considered a pressure outlet, and the symmetry boundary conditions are applied on the two boundaries

in the spanwise direction [48].

5.5 Results and Discussion

The effects of turbulence on the wake recovery process have been previously studied [16, 42, 149]. In

general, higher atmospheric turbulence improves the mixing process and thus accelerates the recovery

process. Figures 5.4, 5.5, and 5.6 show the velocity profiles in the spanwise direction for the LM8.2 blade

profile with different inlet turbulence intensities (TI) at 1, 2, and 6 rotor diameters downstream. It is

seen that the recovery process for turbulence intensity at 10% is significantly faster than at 5%. Thus,

it can be expected that the length of the near wake region shortens as turbulence intensity increases.

Figure 5.4: Velocity profile 1 rotor diameter downstream of the turbine with turbulence intensities of5% and 10%.

To study the influence of turbine blade geometry, a uniform thrust coefficient is used to compare

with the force profile from that of the LM8.2 blade. Figures 5.7, 5.8, and 5.9 show the velocity profiles

for turbulence intensity at 5%, for 1, 2, and 6 rotor diameters downstream of the turbine. At 1 and 2

rotor diameters downstream, the velocity profiles from uniform thrust coefficient do not match well with

results generated from the LM8.2 blade profile. Initially, the profile generated with the LM8.2 blade is

Page 63: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 5. Influence of Rotor Geometry and Atmospheric Turbulence 53

Figure 5.5: Velocity profile 2 rotor diameters downstream of the turbine with turbulence intensities of5% and 10%.

bimodal, and then by 2.5 diameters downstream the velocity profile becomes Gaussian, marking the end

of the near wake region. By 6 rotor diameters downstream, the velocity profiles from both force profiles

become nearly identical. This has strong implications, as it shows that the wake recovery process is

nearly independent of the blade profile in the far wake region. However, if the focus is modelling the

near wake region, which ends at approximately 2.5 rotor diameters downstream of the turbine in this

scenario, then the uniform force approach is not sufficient.

A similar comparison is done to study the effects of increasing atmospheric turbulence on the wake

recovery process under different distribution of forces. Figures 5.10, 5.11, and 5.12 show the velocity

profiles under a turbulence intensity of 10%. At approximately 1.5 rotor diameters downstream, the

velocity profile becomes Gaussian, marking the end of the near wake region. This has implications for

wake models that analytically model or bypass the near wake region, such as the one developed by Jafari

et al. [53], where the near wake region is assumed to be 1 rotor diameter downstream compared to 2.5

and 1.5 rotor diameters for turbulence intensities of 5% and 10%, respectively. Looking at Figures 5.11

and 5.12, the velocity profiles of both LM8.2 and the uniform forces are nearly identical, echoing the

previous results with a 5% turbulence intensity.

Page 64: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 5. Influence of Rotor Geometry and Atmospheric Turbulence 54

Figure 5.6: Velocity profile 6 rotor diameters downstream of the turbine with turbulence intensities of5% and 10%.

5.6 Conclusion

This work set out to answer two main questions: 1) Does the wake recovery depend on the blade profiles?

2) How does turbulence affect the near and far wake regions? Our results show that the near wake

region is strongly affected by the turbine geometry, as evidenced by the shape of the velocity profiles. In

addition, the length of the near wake region is dependent on atmospheric turbulence conditions, where

the near wake region ends 2.5 and 1.5 rotor diameters downstream of the turbine for the turbulence

intensities of 5% and 10%, respectively. Increased turbulence improves the mixing process, shortening

the length of the near wake region. Typically, the effects of the blade geometry become less significant

with higher turbulence intensity.

In the far wake region, increased turbulence allows accelerated recovery as atmospheric turbulence

becomes the dominant factor, which has been reported in previous studies. The results show that in the

far wake region, this recovery is nearly independent of the blade geometry. This suggests that it is not

necessary to model the aerodynamic force distribution and that a uniform force is sufficient to model

the wake in the far wake region. In other words, given the bulk aerodynamic forces that a turbine exerts

on the flow, a uniform force actuator disk is sufficient to model the rotor if the far wake is the region

of interest. For wake models that bypass or analytically model the near wake region, it is necessary to

identify where the near wake region ends and such a transition must be appropriately modelled.

Page 65: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 5. Influence of Rotor Geometry and Atmospheric Turbulence 55

Figure 5.7: Velocity profile 1 rotor diameter downstream of the turbine with different force profiles at aturbulence intensity of 5%.

Figure 5.8: Velocity profile 2 rotor diameters downstream of the turbine with different force profiles ata turbulence intensity of 5%.

Page 66: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 5. Influence of Rotor Geometry and Atmospheric Turbulence 56

Figure 5.9: Velocity profile 6 rotor diameters downstream of the turbine with different force profiles ata turbulence intensity of 5%.

Figure 5.10: Velocity profile 1 rotor diameter downstream of the turbine with different force profiles ata turbulence intensity of 10%.

Page 67: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 5. Influence of Rotor Geometry and Atmospheric Turbulence 57

Figure 5.11: Velocity profile 2 rotor diameters downstream of the turbine with different force profiles ata turbulence intensity of 10%.

Figure 5.12: Velocity profile 6 rotor diameters downstream of the turbine with different force profiles ata turbulence intensity of 10%.

Page 68: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 6

Wake Model for Complex Terrains

6.1 Introduction

Due to cost-competitiveness of onshore wind farms, most of the wind energy development in North

America have been focused on onshore farms [24]. One of the key aspects to wind farm design is the

placement of wind turbines, as wake losses is one of the major losses in wind farms [14]. In order to

optimize turbine placements, wake models are required to evaluate the layout solutions. However, most

of the wake model development have been focused on flat and uniform terrains [18, 19, 20, 21, 22, 23].

Therefore, there is a need to develop wake models capable of simulating wake effects on complex terrains.

During the design optimization process, wake models are used to evaluate the wake effects of different

layout solutions. This process is iterative and requires a large number of solution evaluations, thus low-

cost wake model is necessary to evaluate large pools of candidate solutions. In commercial software,

wake models for complex terrains rely on variations of wake superposition approaches [54]. In addition

to wake superposition, only one wake model capable of simulating wakes on complex terrains has been

developed. This model, virtual particle model [55], utilizes the concept of solving for velocity deficit

instead of velocity by assuming that velocity deficit is analogous to concentration of particles.

The main challenge in developing wake models for complex terrains is reducing the computational

cost. Currently, a number of full-scale computational fluid dynamics (CFD) models [40, 46, 47, 48,

49, 50, 122] have been developed to simulate wake effects. However, these high-fidelity simulations are

costly and difficult to integrate with existing optimization approaches as these models rely on solving the

full Navier-Stokes equations in large computational domains. In this work, the wake development and

propagation process on complex terrains is simulated approximately as an advection-diffusion problem,

reducing its computational cost while maintaining accuracy comparable with full-scale CFD models, and

compatible with the requirements of the wind farm layout optimization problem.

58

Page 69: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 6. Wake Model for Complex Terrains 59

In the following section and subsections, background information on wake turbine wakes is described,

followed by a detailed description of the proposed model and numerical discretization. Finally, the

performance of the proposed model is discussed.

6.1.1 Wind Turbine Wake

There are two main regions in a wind turbine wake, near and far wake regions, shown in Figure 6.1. In

the near wake region, the blade geometry has a strong influence on the velocity and pressure profiles,

while in the far wake region, the velocity profile becomes Gaussian and the blade influence on pressure

gradient is negligible [27, 31, 32].

Figure 6.1: Near and far wake regions of a wind turbine

6.1.2 Actuator Disk Model

The actuator disk model is one of the common methods of modelling wind turbine rotor, owing to

its simplicity. This model assumes an initial wake expansion behind the turbine and its velocity and

pressure profiles are known [51]. The velocity and pressure profiles described by the actuator disk model

are illustrated in Figure 6.2.

Figure 6.2: Actuator disk velocity and pressure distribution

Page 70: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 6. Wake Model for Complex Terrains 60

6.2 Proposed Wake Model

A low-cost numerical wake model that can be used to optimize layouts on complex terrains while main-

taining high accuracy is proposed. It utilizes ideas from the virtual particle model [55] and the turbulent

viscosity development from the Ainslie wake model [15]. The aim of this approach is to make use of

previous development to simplify the problem in order to reduce computational cost.

The model starts with Reynolds-Averege Navier-Stokes equations, shown in Eq.(6.1). Given any

flow over a complex terrain, it must satisfy the Reynolds-averaged Navier-Stoke equations, where u0 is

an unique solution that describes the flow field in a given domain. If a turbine is placed in the same

domain, then the original flow field changes due to the presence of the turbine. The flow field solution

that satisfies the momentum equations can then be defined as uwake, described in Eq.(6.2), where f is

the body force exerted by the turbine. The notation uwake = u0 −∆u is used, where uwake is the wind

velocity due to the presence of a wake, u0 is the wind velocity in the free stream, and ∆u is the velocity

deficit experienced in the flow due to the wake.

u0,j∂u0,i∂xj

= −1

ρ

∂p0∂xi

+ νt,0∂2u0,i∂x2j

(6.1)

uwake,j∂uwake,i∂xj

= −1

ρ

∂pwake∂xi

+ νt,wake∂2uwake,i∂x2j

+ fi (6.2)

In order to obtain the difference (velocity deficit) between the two flow fields, Eq.6.1) is subtracted

by Eq.(6.2),

u0,j∂u0,i∂xj

− uwake,j∂uwake,i∂xj

= −1

ρ

(∂p0∂xi− ∂pwake

∂xi

)+ νt,0

∂2u0,i∂x2j

− νt,wake∂2uwake,i∂x2j

− fi. (6.3)

If Eq.(6.3) can be solved, then the influence of a turbine on the flow field can be quantified. Before

this Eq.(6.3) can be solved, a number of simplifications and assumptions can be applied to reduce the

complexity.

Firstly, for the purpose of wind farm layout design, turbines are placed in the far wake region of

upstream turbines, where pressure gradient caused by the turbine is considered to be negligible, and

the turbulent viscosity is assumed to be approximately the same (νt,wake ≈ νt,0). Furthermore, to avoid

directly simulating the near wake region, where the flow field is dependent on turbine blade profiles, the

body force term f is then described using the actuator disk model, by prescribing the wake expansion

and velocity profile at the end of the near wake. Then, uwake = u0−∆u can be substituted into Eq.(6.3),

which simplifies to

Page 71: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 6. Wake Model for Complex Terrains 61

(u0,j −∆uj)∂∆ui∂xj

+ ∆uj∂u0,i∂xj

= νt,wake∂2∆ui∂x2j

. (6.4)

The first term of the equation represents the advection term, describing how the velocity deficit is

advected. The second term accounts for the effects of velocity gradients while the third term describes

the diffusion of velocity deficit. Furthermore, the advection term can be assumed to be more dominant

than the diffusion term in the streamwise direction.

Figure 6.3: Actuator disk theory, the wake experiences an initial wake expansion in the near wake.

Equation (6.4) describes the influence of a turbine on the flow field. The variable of interest is

velocity deficit, ∆u. The free stream velocity u0 and viscosity νt,wake are obtained from full-scale CFD

or experimental measurements. As discussed previously, the actuator disk model is used to describe the

momentum extracted due to the presence of a turbine. It states that the wake experiences an initial

wake expansion (Fig. 6.3) [51, 53] in the near wake and this expansion at the end of the near wake (≈

2 rotor diameters downstream of the turbine) is described as,

rwake,2D =rturbine(1− a)

1− 2a(6.5)

and

uwake,2D = u0 − 2a, (6.6)

respectively. In these equations, rturbine is the rotor radius of the turbine, a = 1−√1−CT2 is the induction

factor and CT is the thrust coefficient of the turbine. Due to the presence of the turbine, a modification

to the turbulent viscosity is required to describe the non-equilibrium between the velocity field and

the turbulence field [15], up to about 5 diameters downstream of the rotor. Ainslie described this

non-equilibrium in the form of,

νt,wake(x) = νt,c + νt,0 = F [2k1rwake,2D(u0(x)− uc(x))] + νt,0(x) (6.7)

Page 72: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 6. Wake Model for Complex Terrains 62

F =

0.65 +

(x−4.523.32

) 13 2 ≤ x ≤ 5.5

1 x > 5.5

(6.8)

where νt,c and uc are the turbulent viscosity and velocity at the wake centerline, respectively, and x is the

distance downstream of the turbine in rotor diameter. The constant k1 is found experimentally. The term

u0(x) − uc(x), defines the velocity deficit in the wake centerline. The filter function F (x) accounts for

the non-equilibrium based on fitting with experimental data. Ainslie’s νt,c function produced inadequate

fit when compared our full-scale CFD simulation, thus we redefine νt,c to ensure better fit with full-scale

CFD simulation data, discussed in the next section.

6.3 Results and Discussion

In this section, the turbulent viscosity term, νt,c, is redefined to fit better with full-scale CFD simulation

results on a flat terrain. Then, the model is compared with full-scale CFD simulations of a turbine

placed on a complex terrain found at the Gros-Morne Wind Farm in Quebec.

(a) Velocity contour for CFD simulation.

(b) Velocity contour for proposed model.

Figure 6.4: Streamwise velocity contour for both CFD and proposed model.

Page 73: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 6. Wake Model for Complex Terrains 63

6.3.1 Turbulent Viscosity Model Fitting

A turbine of 80 m in both rotor diameter and hub height is placed in a 3000 m x 1000 m x 1000

m domain. The turbine is placed 1000 m downstream (x-direction) of the inlet, 80 m for hub height

(y-direction), and 500 m in the spanwise direction (z-direction). At the inlet, a logarithmic velocity

profile is prescribed, u(y) = u∗

κ ln( yy0 ), where u(y) represents the streamwise wind speed with varying

height y and u∗ is the friction velocity which is assumed to be 0.4 m/s [150]. Turbulent kinetic energy

and dissipation rate at the inlet are prescribed as k = (u∗)2√Cµ

and ε(y) = (u∗)3

κy , where Cµ = 0.033 and

κ = 0.4. At the top boundary, velocity and turbulence quantities are correspondingly prescribed at 1000

m. The side boundaries are set as symmetry boundary conditions and a pressure outlet is prescribed at

the outlet boundary. The turbulence model used for full-scale CFD simulations is Reynolds stress model

(RSM), with Cµ = 0.033, identical to that used by Makridis and Chick [48].

As previously mentioned, the νt,c(x) from Eq.(6.7) is redefined as νt′(x, y, z) and fitted to full-

scale CFD simulations to ensure closer comparison between the proposed model and full-scale CFD

simulations. As observed by Ainslie, νt′ is a function of velocity deficit, and that this relationship is

approximately parabolic, shown in Figure 6.5. The fitted function is shown in Eq.(6.9),

νt′(x, y, z) = −7.28916× 10−5(2rwake,2D[u0(x, y, z)− uwake(x, y, z)])2

+ 3.64535× 10−2(2rwake,2D[u0(x, y, z)− uwake(x, y, z)]) + 1.6457.(6.9)

In the proposed model, the domain is discretized evenly into 10 m x 10 m x 10 m cells, or 3 million

cells in this domain. Comparatively, the full-scale CFD simulation uses 0.78 million cells. This is due to

the advanced meshing capabilities available in commercial CFD packages, while only uniform-sized cells

can be assigned in the in-house code for the proposed model. Then the velocity profiles of the full-scale

CFD and the proposed models at 5 and 10 rotor diameters downstream of the rotor are compared, shown

in Figures 6.6a and 6.6b. Looking at these profiles, there is a very good agreement between the two

models in terms of accuracy with error less than 5% and 3% for 5 and 10 rotor diameters, respectively;

however, the full-scale CFD simulation took approximately 3 hours to complete with 3 processors while

the proposed model converged in 2 minutes with a single processor. These wake effects are apparent

up to 20 rotor diameters, and the error is less than 4%. Based on these results, the proposed model

is able to describe the wake recovery process accurately while keeping the computational cost orders of

magnitude lower.

Page 74: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 6. Wake Model for Complex Terrains 64

Figure 6.5: Increase in turbulence viscosity inside a turbine wake.

6.3.2 Complex Terrains

In this test case, the same turbine at the flat terrain scenario is placed on a complex terrain based on

the Gros-Morne Wind Farm in Quebec, Canada, shown in Figure 6.7. The inlet profile is prescribed

using the log law previously described, with a surface roughness length, y0, of 0.0002 m, corresponding

to open sea terrains, as the wind blows inland from the sea. As with the flat terrain simulation, the

friction velocity is assumed to be 0.4 m/s [150]. The roughness length of domain is prescribed to be 0.04

m, corresponding to roughness of crop terrain [51].

The simulation domain is 4500 m in the streamwise direction, 1790 m in the spanwise direction, and

1500 m in height. The full-scale CFD domain is discretized into 2.3 million cells, with higher resolution

closer to the ground. In the proposed model, the domain is discretized into 10 m x 10 m x 10 m cells,

with a total of approximately 12 million cells in the domain.

Initially, a full-scale CFD simulation without the turbine is conducted to obtain the natural flow

field over the complex terrain. Then, using the actuator disk model, the velocity profile is prescribed

2 rotor diameters downstream of the turbine. Due to the terrain topography, the velocity has multiple

components, thus the components of the velocity deficit prescribed at the end of the near wake should

correspond to the velocity components at that location. In other words, the velocity deficit is prescribed

in the direction of the velocity vector two diameters downstream.

Page 75: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 6. Wake Model for Complex Terrains 65

(a) Velocity profile at 5 rotor diameters down-stream of turbine.

(b) Velocity profile at 5 rotor diameters down-stream of turbine.

Figure 6.6: Streamwise velocity profiles on a flat terrain with new turbulent viscosity fitting.

Figure 6.8 shows the streamwise velocity at hub height for both full-scale CFD and proposed models,

and Figures 6.9a and 6.9b show the cross-sectional streamwise velocity contour for both models. Overall,

the proposed model matches the full-scale CFD results very well (less than 3.5% error at hub height)

but using only a small fraction of the computational cost. The simulation times of the full-scale CFD

model and the proposed model are 11 hours with 3 processors and 9 minutes with a single processor,

respectively. This low-cost and accurate wake model will drastically speed-up the design optimization

process of wind farms located on complex terrains.

6.4 Conclusions

This work focuses on developing a low cost and accurate wake model capable of simulating wake effects

on complex terrains. The model solves a simplified variation of the Navier-Stokes equations with three

major simplifications to reduce computational cost while maintaining accuracy:

1. Near wake is modelled using the actuator disk theory

2. Far wake is the region of interest

3. Velocity deficit is the variable of interest

This model was validated against CFD simulation of a turbine placed on the terrain of Gros-Morne Wind

Farm in Quebec. The proposed model allows for fast simulation of wake effects on complex terrains,

making it ideal for designing wind farm layouts on complex terrains.

Page 76: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 6. Wake Model for Complex Terrains 66

Figure 6.7: CFD domain with terrain based on Gros-Morne Wind Farm.

Figure 6.8: Streamwise velocity at hub height.

Page 77: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 6. Wake Model for Complex Terrains 67

(a) CFD simulation.

(b) Proposed model.

Figure 6.9: Streamwise velocity on a complex terrain using CFD and proposed model.

Page 78: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 7

Future Work

This thesis presents a new optimization algorithm, a wake model, and a wake interaction model to

tackle the challenges of designing the layout of wind farms on complex terrains, as well as a study

for quantifying the influence of atmospheric turbulence and terrain on wake propagation and recovery

process. While the methodologies used to develop these models and algorithm are fundamentally sound,

the predicted improvements cannot be verified through real wind farm data. In addition, validating

CFD simulations with field data remains difficult due to the experimental uncertainties. One of the

major assumptions made in this work is the condition of atmospheric stability, which is considered to be

constant and neutral. The effects of atmospheric stability and other physical phenomena such as wake

meandering on wind farm design need to be captured in future work. In terms of future research, there

are several possible directions.

Wake Modelling and Optimization Integration

In this thesis, we proposed an optimization algorithm that integrates full-scale CFD simulations into the

optimization process that reduces computational cost, as well as a fast wake model for complex terrains

with similar accuracy as full-scale CFD simulations. This wake model can be used to replace full-scale

CFD simulations to cut optimization times by orders of magnitude. This is a very promising method

for wind farm layout design.

Robust Wind Farm Design

Uncertainty is present in all stages of engineering design. The operations research community is very

familiar with optimization under uncertainty. In terms of wind farm layout optimization, Zhang et al.

[151] included wind direction uncertainty into the layout optimization process. The concept of robust

design can also be applied to uncertainty in wake modelling. A robust wind farm design will help reduce

68

Page 79: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Chapter 7. Future Work 69

fluctuations in power generation due to wake losses [152].

Real-Time Optimization of Wind Farm Production

After a farm has been designed and installed in place. It is likely that the energy production does not

match that of forecast found through simulations. We observed in the literature that many wake models

underestimate the wake effects [28] experienced in the field. Wind farm control is an active research

area that improves farm performance by controlling individual turbines. A recent study by Fleming et

al. [153] shows that pitch and yaw angles can induce deflections in the wake, reducing the wake losses

experienced by downstream turbines. Real-time optimization is possible with the tools that we have

developed, and as such, is a very promising research direction.

Hybrid Renewable Energy Systems

Wind energy is typically not a stand-alone source of energy, due to its intermittent nature. We have

experienced proliferation in energy storage research in recent years. Particularly for wind energy, pumped

hydro [154], flywheels [155], and smart battery technologies [156] are promising storage technologies that

can stabilize the intermittency. Furthermore, wind energy can be used in conjunction with other energy

sources such as solar and gas turbines. The integration of these systems to provide cost-effective and

stable power will become increasingly important.

Page 80: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

Bibliography

[1] R. R. McKitrick, “Environmental and economic consequences of Ontarios Green Energy Act,”

tech. rep., Fraser Institute, 2013.

[2] “Wind energy in Canada: Realizing the opportunity,” tech. rep., KPMG, 2013.

[3] Vaisala 3TIER-Services, “Free wind and solar resource maps,” 2015.

[4] C. B. Hasager, L. Rasmussen, A. Pena, L. E. Jensen, and P.-E. Rethore, “Wind farm wake: The

Horns Rev photo case,” Energies, vol. 6, no. 2, pp. 696–716, 2013.

[5] N. Jensen, “A note on wind generator interaction,” tech. rep., Riso National Laboratory, Denmark,

1983.

[6] S. Frandsen, R. Barthelmie, S. Pryor, O. Rathmann, S. Larsen, J. Hjstrup, and M. Thgersen,

“Analytical modelling of wind speed deficit in large offshore wind farms,” Wind Energy, vol. 9,

no. 1-2, pp. 39–53, 2006.

[7] A. El Kasmi and C. Masson, “An extended k–ε model for turbulent flow through horizontal-

axis wind turbines,” Journal of Wind Engineering and Industrial Aerodynamics, vol. 96, no. 1,

pp. 103–122, 2008.

[8] Environment Canada, “Canadian wind energy atlas,” 2011.

[9] S. Shukla and S. Sawyer, “30 years of policies for wind energy,” tech. rep., The International

Renewable Energy Agency (IRENA), 01 2012.

[10] European Wind Energy Association, “Wind energy scenarios for 2030,” 2015.

[11] Bloomberg New energy Finance, “Sustainable energy in America: 2016 factbook,” 2016.

[12] Canadian Wind Energy Association, “Wind energy continues rapid growth in Canada in 2015,”

2016.

70

Page 81: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

BIBLIOGRAPHY 71

[13] X. Lu, M. B. McElroy, and J. Kiviluoma, “Global potential for wind-generated electricity,” Pro-

ceedings of the National Academy of Sciences, vol. 106, no. 27, pp. 10933–10938, 2009.

[14] R. J. Barthelmie, K. Hansen, S. T. Frandsen, O. Rathmann, J. G. Schepers, W. Schlez, J. Phillips,

K. Rados, A. Zervos, E. S. Politis, and P. K. Chaviaropoulos, “Modelling and measuring flow and

wind turbine wakes in large wind farms offshore,” Wind Energy, vol. 12, no. 5, pp. 431–444, 2009.

[15] J. Ainslie, “Calculating the flowfield in the wake of wind turbines,” Journal of Wind Engineering

and Industrial Aerodynamics, vol. 27, no. 13, pp. 213 – 224, 1988.

[16] D. J. Renkema, “Validation of wind turbine wake models,” Master’s thesis, Delft University of

Technology, 2007.

[17] R. J. Barthelmie and L. Jensen, “Evaluation of wind farm efficiency and wind turbine wakes at

the Nysted offshore wind farm,” Wind Energy, vol. 13, no. 6, pp. 573–586, 2010.

[18] P. Y. Zhang, “Topics in wind farm layout optimization: Analytical wake models, noise propagation,

and energy production,” Master’s thesis, University of Toronto, 2013.

[19] S. Turner, D. Romero, P. Zhang, C. Amon, and T. Chan, “A new mathematical programming

approach to optimize wind farm layouts,” Renewable Energy, vol. 63, pp. 674–680, 2014.

[20] G. Mosetti, C. Poloni, and B. Diviacco, “Optimization of wind turbine positioning in large wind-

farms by means of a genetic algorithm,” Journal of Wind Engineering and Industrial Aerodynamics,

vol. 51, no. 1, pp. 105–116, 1994.

[21] S. Grady, M. Hussaini, and M. Abdullah, “Placement of wind turbines using genetic algorithms,”

Renewable Energy, vol. 30, no. 2, pp. 259–270, 2005.

[22] J. Park and K. H. Law, “Layout optimization for maximizing wind farm power production using

sequential convex programming,” Applied Energy, vol. 151, pp. 320 – 334, 2015.

[23] X. Gao, H. Yang, and L. Lu, “Study on offshore wind power potential and wind farm optimization

in Hong Kong,” Applied Energy, vol. 130, pp. 519 – 531, 2014.

[24] “2015 U.S. wind industry market reports,” tech. rep., AWEA, 2015.

[25] W. Tong, S. Chowdhury, J. Zhang, and A. Messac, “Impact of different wake models on the

estimation of wind farm power generation,” in AIAA/ISSMO 2012 Multidisciplinary Analysis and

Optimization Conference 2012. Indianapolis, IN, USA, (Indianapolis, IN, USA), September.

Page 82: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

BIBLIOGRAPHY 72

[26] J. Cleve, M. Greiner, P. Enevoldsen, B. Birkemose, and L. Jensen, “Model-based analysis of wake-

flow data in the Nysted offshore wind farm,” Wind Energy, vol. 12, no. 2, pp. 125–135, 2009.

[27] B. Sanderse, “Aerodynamics of wind turbine wakes,” Energy Research Center of the Netherlands

(ECN), ECN-E–09-016, Petten, The Netherlands, Tech. Rep, 2009.

[28] R. Shakoor, M. Y. Hassan, A. Raheem, and Y.-K. Wu, “Wake effect modeling: A review of

wind farm layout optimization using Jensen’s model,” Renewable and Sustainable Energy Reviews,

vol. 58, pp. 1048–1059, 2016.

[29] A. Crespo, J. Hernandez, and S. Frandsen, “Survey of modelling methods for wind turbine wakes

and wind farms,” Wind energy, vol. 2, no. 1, pp. 1–24, 1999.

[30] J. G. Schepers, ENDOW: Validation and improvement of ECNs wake model. Technical Report

ECN-C-03-034. Energy research Centre of the Netherlands, 2003.

[31] P. H. Tari, Near-Wake Flow Dynamics of a Horizontal Axis Wind Turbine. PhD thesis, The

University of Western Ontario, 2012.

[32] M.-K. Liu, M. Yocke, and T. Myers, “Mathematical model for the analysis of wind-turbine wakes,”

Journal of Energy, vol. 7, no. 1, pp. 73–78, 1983.

[33] S. Yamani Douzi Sorkhabi, D. A. Romero, G. K. Yan, M. D. GU, J. Moran, M. Morgenroth, and

C. H. Amon, “Constrained multi-objective wind farm layout optimization: Introducing a novel

constraint handling approach based on constraint programming,” in Proceedings of the ASME

2015 International Design Engineering Technical Conferences and Computers & Information in

Engineering Conference, Boston, MA, USA, 2015.

[34] T. Ishihara, A. Yamaguchi, and Y. Fujino, “Development of a new wake model based on a wind

tunnel experiment,” Global wind power, p. 1, 2004.

[35] G. C. Larsen, “A SimpleWake Calculation Procedure,” tech. rep., Riso National Laboratory, 1988.

[36] “Wind farm design software: theory manual,” tech. rep., Garrad Hassan and Partners Ltd., 2011.

[37] M. Brower and N. Robinson, “The open wind deep-array wake model: development and valida-

tion,” tech. rep., AWS Truepower, 2010.

[38] B. Lange, H.-P. Waldl, A. G. Guerrero, D. Heinemann, and R. J. Barthelmie, “Modelling of offshore

wind turbine wakes with the wind farm program FLaP,” Wind Energy, vol. 6, no. 1, pp. 87–104,

2003.

Page 83: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

BIBLIOGRAPHY 73

[39] J. Schepers, ENDOW: Validation and improvement of ECN’s wake model. Energy research Centre

of the Netherlands ECN, 2003.

[40] J. Prospathopoulos, E. Politis, K. Rados, and P. Chaviaropoulos, “Evaluation of the effects of

turbulence model enhancements on wind turbine wake predictions,” Wind Energy, vol. 14, no. 2,

pp. 285–300, 2011.

[41] A. El Kasmi and C. Masson, “An extended k–ε model for turbulent flow through horizontal-

axis wind turbines,” Journal of Wind Engineering and Industrial Aerodynamics, vol. 96, no. 1,

pp. 103–122, 2008.

[42] P.-E. Rethore, “Wind turbine wake in atmospheric turbulence,” 2009.

[43] N. Troldborg, “Actuator line modeling of wind turbine wakes,” 2009.

[44] D. Sturge, D. Sobotta, R. Howell, A. While, and J. Lou, “A hybrid actuator disc–full rotor CFD

methodology for modelling the effects of wind turbine wake interactions on performance,” Renew-

able Energy, vol. 80, pp. 525–537, 2015.

[45] J.-O. Mo, A. Choudhry, M. Arjomandi, and Y.-H. Lee, “Large eddy simulation of the wind tur-

bine wake characteristics in the numerical wind tunnel model,” Journal of Wind Engineering and

Industrial Aerodynamics, vol. 112, pp. 11–24, 2013.

[46] A. E. Kasmi and C. Masson, “An extended model for turbulent flow through horizontal-axis wind

turbines,” Journal of Wind Engineering and Industrial Aerodynamics, vol. 96, no. 1, pp. 103 –

122, 2008.

[47] Y.-T. Wu and F. Porte-Agel, “Atmospheric turbulence effects on wind-turbine wakes: An LES

study,” Energies, vol. 5, no. 12, pp. 5340–5362, 2012.

[48] A. Makridis and J. Chick, “Validation of a CFD model of wind turbine wakes with terrain effects,”

Journal of Wind Engineering and Industrial Aerodynamics, vol. 123, pp. 12–29, 2013.

[49] J. Schmidt and B. Stoevesandt, “Modelling complex terrain effects for wind farm layout optimiza-

tion,” in Journal of Physics: Conference Series, vol. 524, p. 012136, IOP Publishing, 2014.

[50] J. Schmidt and B. Stoevesandt, “Wind farm layout optimization in complex terrain with CFD

wakes,” 2015.

[51] J. F. Manwell, J. G. McGowan, and A. L. Rogers, Wind energy explained: theory, design and

application. John Wiley & Sons, 2010.

Page 84: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

BIBLIOGRAPHY 74

[52] P.-E. Rethore, P. Laan, N. Troldborg, F. Zahle, and N. N. Sørensen, “Verification and validation

of an actuator disc model,” Wind Energy, vol. 17, no. 6, pp. 919–937, 2014.

[53] S. Jafari, N. Chokani, and R. Abhari, “Simulation of wake interactions in wind farms using an

immersed wind turbine model,” Journal of Turbomachinery, vol. 136, no. 6, p. 061018, 2014.

[54] J. Feng and W. Z. Shen, “Wind farm layout optimization in complex terrain: A preliminary study

on a Gaussian hill,” in Journal of Physics: Conference Series, vol. 524, p. 012146, IOP Publishing,

2014.

[55] M. Song, K. Chen, Z. He, and X. Zhang, “Wake flow model of wind turbine using particle simu-

lation,” Renewable energy, vol. 41, pp. 185–190, 2012.

[56] J. F. Herbert-Acero, O. Probst, P.-E. Rethore, G. C. Larsen, and K. K. Castillo-Villar, “A review

of methodological approaches for the design and optimization of wind farms,” Energies, vol. 7,

no. 11, pp. 6930–7016, 2014.

[57] G. K. Hartwig Dobesch, Basic meteorological concepts and recommendations for the exploitation of

wind energy in the atmospheric boundary layer. Central Institute for Meteorology and Geodynamics

(ZAMG), 2001.

[58] H. M. Erik Djerf, “Evaluation of the software program Windfarm and comparisons with measured

data from Alsvik,” tech. rep., Aeronautical Research Institute, 2000.

[59] S. Donovan, “An improved mixed integer programming model for wind farm layout optimisation,”

in Proceedings of the 41th Annual Conference of the Operations Research Society. Wellington, New

Zealand, 2006.

[60] A. Mittal, Optimization of the Layout of Large Wind Farms using a Genetic Algorithm. PhD

thesis, Case Western Reserve University, 2010.

[61] A. Kusiak and Z. Song, “Design of wind farm layout for maximum wind energy capture,” Renewable

Energy, vol. 35, no. 3, pp. 685–694, 2010.

[62] B. L. Du Pont and J. Cagan, “An extended pattern search approach to wind farm layout opti-

mization,” Journal of Mechanical Design, vol. 134, p. 081002, 2012.

[63] S. Chowdhury, J. Zhang, A. Messac, and L. Castillo, “Unrestricted wind farm layout optimiza-

tion (UWFLO): Investigating key factors influencing the maximum power generation,” Renewable

Energy, vol. 38, no. 1, pp. 16–30, 2012.

Page 85: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

BIBLIOGRAPHY 75

[64] J. Serrano Gonzalez, M. Burgos Payan, and J. Riquelme Santos, “Wind farm optimal design

including risk,” in Modern Electric Power Systems (MEPS), 2010 Proceedings of the International

Symposium, pp. 1–6, IEEE.

[65] W. Y. Kwong, P. Y. Zhang, D. Romero, J. Moran, M. Morgenroth, and C. Amon, “Multi-objective

wind farm layout optimization considering energy generation and noise propagation with NSGA-

II,” Journal of Mechanical Design, vol. 136, no. 9, p. 091010, 2014.

[66] L. Chen and E. MacDonald, “Considering landowner participation in wind farm layout optimiza-

tion,” Journal of Mechanical Design, vol. 134, p. 084506, 2012.

[67] L. Chen and E. MacDonald, “A system-level cost-of-energy wind farm layout optimization with

landowner modeling,” Energy Conversion and Management, vol. 77, pp. 484–494, 2014.

[68] S. Chowdhury, J. Zhang, A. Messac, and L. Castillo, “Optimizing the arrangement and the selection

of turbines for wind farms subject to varying wind conditions,” Renewable Energy, vol. 52, pp. 273–

282, 2013.

[69] P. Y. Zhang, D. A. Romero, J. C. Beck, and C. H. Amon, “Solving wind farm layout optimiza-

tion with mixed integer programs and constraint programs,” EURO Journal on Computational

Optimization, vol. 2, no. 3, pp. 195–219, 2014.

[70] P. Fagerfjall, “Optimizing wind farm layout–more bang for the buck using mixed integer linear

programming,” Master’s thesis, Chalmers University of Technology and Gothenburg University,

2010.

[71] B. Perez, R. Mınguez, and R. Guanche, “Offshore wind farm layout optimization using mathemat-

ical programming techniques,” Renewable Energy, vol. 53, pp. 389–399, 2013.

[72] N. Quan and H. Kim, “A mixed integer linear programming formulation for unrestricted wind

farm layout optimization,” in ASME 2015 International Design Engineering Technical Conferences

and Computers and Information in Engineering Conference, pp. V02AT03A028–V02AT03A028,

American Society of Mechanical Engineers, 2015.

[73] S. Donovan, G. Nates, H. Waterer, and R. Archer, “Mixed integer programming models for wind

farm design,” in Slides used at MIP 2008 Workshop on Mixed Integer Programming, Columbia

University, New York City, 2008.

[74] R. Archer, G. Nates, S. Donovan, and H. Waterer, “Wind turbine interference in a wind farm

layout optimization mixed integer linear programming model,” Wind Engineering, vol. 35, no. 2,

pp. 165–175, 2011.

Page 86: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

BIBLIOGRAPHY 76

[75] S. Chowdhury, J. Zhang, A. Messac, and L. Castillo, “Characterizing the influence of land config-

uration on the optimal wind farm performance,” in ASME 2011 International Design Engineering

Technical Conferences & Computers and Information in Engineering Conference (IDETC/CIE

2011), No. DETC2011-48731, ASME. Washington, DC, USA, 2011.

[76] S. Lu and H. Kim, “Wind farm layout design optimization through multidisciplinary design opti-

mization with complementarity constraints,” in 12th AIAA Aviation Technology, Integration, and

Operations (ATIO) Conference and 14th AIAA/ISSMO Multidisciplinary Analysis and Optimiza-

tion Conference, p. 5429, 2012.

[77] M. Wagner, J. Day, and F. Neumann, “A fast and effective local search algorithm for optimizing

the placement of wind turbines,” Renewable Energy, vol. 51, pp. 64–70, 2013.

[78] S. Y. D. Sorkhabi, D. A. Romero, G. K. Yan, M. D. Gu, J. Moran, M. Morgenroth, and C. H.

Amon, “The impact of land use constraints in multi-objective energy-noise wind farm layout

optimization,” Renewable Energy, vol. 85, pp. 359–370, 2016.

[79] J. S. Gonzalez, A. G. Gonzalez Rodriguez, J. C. Mora, J. R. Santos, and M. B. Payan, “Optimiza-

tion of wind farm turbines layout using an evolutive algorithm,” Renewable Energy, vol. 35, no. 8,

pp. 1671–1681, 2010.

[80] J. S. Gonzalez, M. Burgos Payan, and J. M. Riquelme-Santos, “Optimization of wind farm turbine

layout including decision making under risk,” Systems Journal, IEEE, vol. 6, no. 1, pp. 94–102,

2012.

[81] B. Saavedra-Moreno, S. Salcedo-Sanz, A. Paniagua-Tineo, L. Prieto, and A. Portilla-Figueras,

“Seeding evolutionary algorithms with heuristics for optimal wind turbines positioning in wind

farms,” Renewable Energy, vol. 36, no. 11, pp. 2838–2844, 2011.

[82] R. Archer, G. Nates, S. Donovan, and H. Waterer, “Wind turbine interference in a wind farm layout

optimization mixed integer linear programming model,” Wind Engineering, vol. 35, pp. 165–175,

Apr. 2011.

[83] M. Bilbao and E. Alba, “Simulated annealing for optimization of wind farm annual profit,” in

2009 2nd International Symposium on Logistics and Industrial Informatics, vol. 0, pp. 1–5, IEEE,

Sept. 2009.

[84] M. Bilbao and E. Alba, “GA and PSO applied to wind energy optimization,” in CACIC Conference

Proceedings, (Jujuy, Argentina), 2009.

Page 87: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

BIBLIOGRAPHY 77

[85] J. Castro Mora, J. M. Calero Baron, J. M. Riquelme Santos, and M. Burgos Payan, “An evolutive

algorithm for wind farm optimal design,” Neurocomputing, vol. 70, pp. 2651–2658, Oct. 2007.

[86] C. N. Elkinton, J. F. Manwell, and J. G. McGowan, “Offshore wind farm layout optimization

(OWFLO) project : Preliminary results,” in Collection of Technical Papers - 44th AIAA Aerospace

Sciences Meeting, pp. 11964–11972, 2006.

[87] C. N. Elkinton, J. F. Manwell, and J. G. McGowan, “Algorithms for offshore wind farm layout

optimization,” Wind Engineering, vol. 32, pp. 67–84, Oct. 2008.

[88] A. Emami and P. Noghreh, “New approach on optimization in placement of wind turbines within

wind farm by genetic algorithms,” Renewable Energy, vol. 35, pp. 1559–1564, July 2010.

[89] J. S. Gonzalez, A. G. Rodrıguez, J. C. Mora, M. Burgos Payan, and J. M. R. Santos, “Overall

design optimization of wind farms,” Renewable Energy, vol. 36, pp. 1973–1982, July 2011.

[90] H. S. Huang, “Distributed Genetic Algorithm for Optimization of Wind Farm Annual Profits,”

in 2007 International Conference on Intelligent Systems Applications to Power Systems, pp. 1–6,

Ieee, Nov. 2007.

[91] H.-s. Huang, “Efficient hybrid distributed genetic algorithms for wind turbine positioning in large

wind farms,” 2009 IEEE International Symposium on Industrial Electronics, pp. 2196–2201, July

2009.

[92] J. Meyers and C. Meneveau, “Optimal turbine spacing in fully developed wind farm boundary

layers,” Wind Energy, vol. 15, pp. 305–317, Mar. 2012.

[93] M. Samorani, “The Wind Farm Layout Optimization Problem,” Optimization, pp. 1–18.

[94] M. Wagner, F. Neumann, and U.-m. O’Reilly, “Optimizing the layout of 1000 wind turbines,” in

European Wind Energy Association Annual Event, 2011.

[95] K. Chen, M. Song, X. Zhang, and S. Wang, “Wind turbine layout optimization with multiple hub

height wind turbines using greedy algorithm,” Renewable Energy, vol. 96, pp. 676–686, 2016.

[96] D. Luckehe, M. Wagner, and O. Kramer, “On evolutionary approaches to wind turbine place-

ment with geo-constraints,” in Proceedings of the 2015 on Genetic and Evolutionary Computation

Conference, pp. 1223–1230, ACM, 2015.

[97] C. Szafron, “Offshore windfarm layout optimization,” 2010 9th International Conference on En-

vironment and Electrical Engineering, pp. 542–545, 2010.

Page 88: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

BIBLIOGRAPHY 78

[98] S. Sisbot, O. Turgut, M. Tunc, and U. Camdal, “Optimal positioning of wind turbines on Gokceada

using multi-objective genetic algorithm,” Wind Energy, vol. 13, pp. 297–306, May 2010.

[99] P.-E. Rethore, “State of the art in wind farm layout optimization,” 2010.

[100] P.-e. Rethore, P. Fuglsang, G. C. Larsen, T. Buhl, T. J. Larsen, and H. A. Madsen, “TopFarm:

Multi-fidelity optimization of offshore wind farm,” in Proceedings of the Twenty-first (2011) Inter-

national Offshore and Polar Engineering Conference, no. 2007, (Maui, Hawaii), 2011.

[101] G. Marmidis, S. Lazarou, and E. Pyrgioti, “Optimal placement of wind turbines in a wind park

using Monte Carlo simulation,” Renewable Energy, vol. 33, pp. 1455–1460, July 2008.

[102] R. Shakoor, M. Y. Hassan, A. Raheem, and N. Rasheed, “Wind farm layout optimization using

area dimensions and definite point selection techniques,” Renewable Energy, vol. 88, pp. 154–163,

2016.

[103] I. Mustakerov and D. Borissova, “Wind turbines type and number choice using combinatorial

optimization,” Renewable Energy, vol. 35, pp. 1887–1894, Sept. 2010.

[104] K. Chen, M. Song, and X. Zhang, “The investigation of tower height matching optimization for

wind turbine positioning in the wind farm,” Journal of Wind Engineering and Industrial Aerody-

namics, vol. 114, pp. 83–95, 2013.

[105] K. Chen, M. Song, and X. Zhang, “The iteration method for tower height matching in wind farm

design,” Journal of Wind Engineering and Industrial Aerodynamics, vol. 132, pp. 37–48, 2014.

[106] M. Iqbal, M. Azam, M. Naeem, A. Khwaja, and A. Anpalagan, “Optimization classification,

algorithms and tools for renewable energy: A review,” Renewable and Sustainable Energy Reviews,

vol. 39, pp. 640–654, 2014.

[107] M. Song, K. Chen, X. Zhang, and J. Wang, “The lazy greedy algorithm for power optimization of

wind turbine positioning on complex terrain,” Energy, vol. 80, pp. 567–574, 2015.

[108] M. Song, K. Chen, Z. He, and X. Zhang, “Bionic optimization for micro-siting of wind farm on

complex terrain,” Renewable Energy, vol. 50, pp. 551–557, 2013.

[109] M. Song, K. Chen, Z. He, and X. Zhang, “Optimization of wind farm micro-siting for complex

terrain using greedy algorithm,” Energy, vol. 67, pp. 454–459, 2014.

[110] C. Wan, J. Wang, G. Yang, and X. Zhang, “Optimal siting of wind turbines using real coded genetic

algorithms,” in Proceedings of European Wind Energy Association Conference and Exhibition.

Stockholm, Sweden, 2009.

Page 89: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

BIBLIOGRAPHY 79

[111] U. Aytun Ozturk and B. A. Norman, “Heuristic methods for wind energy conversion system

positioning,” Electric Power Systems Research, vol. 70, no. 3, pp. 179–185, 2004.

[112] A. Schrijver, Theory of linear and integer programming. John Wiley & Sons, 1998.

[113] R. J. Barthelmie, S. Pryor, S. T. Frandsen, K. S. Hansen, J. Schepers, K. Rados, W. Schlez,

A. Neubert, L. Jensen, and S. Neckelmann, “Quantifying the impact of wind turbine wakes on

power output at offshore wind farms,” Journal of Atmospheric and Oceanic Technology, vol. 27,

no. 8, pp. 1302–1317, 2010.

[114] F. M. White, “Fluid mechanics, 5th edition.,” Boston: McGraw-Hill Book Company, 2003.

[115] G. G. Wang and S. Shan, “Review of metamodeling techniques in support of engineering design

optimization,” Journal of Mechanical Design, vol. 129, no. 4, pp. 370–380, 2007.

[116] D. A. Romero, C. H. Amon, and S. Finger, “Multiresponse metamodeling in simulation-based

design applications,” Journal of Mechanical Design, vol. 134, no. 9, p. 091001, 2012.

[117] R. Moore, C. J. Paredis, and D. A. Romero, “Value-based global optimization,” Journal of Me-

chanical Design, vol. 136, pp. 1–14, 2014.

[118] K. S. Hansen, “Wake measurements used in the model evaluation,” tech. rep., Technical University

of Denmark, 2008.

[119] S. Donovan, “Wind farm optimization,” in Proceedings of the 40th Annual ORSNZ Conference,

pp. 196–205, 2005.

[120] W. J. Gutjahr, “ACO algorithms with guaranteed convergence to the optimal solution,” Informa-

tion processing letters, vol. 82, no. 3, pp. 145–153, 2002.

[121] R. M. Karp, Reducibility among combinatorial problems. Springer, 1972.

[122] R. Barthelmie and S. Pryor, “An overview of data for wake model evaluation in the Virtual Wakes

Laboratory,” Applied Energy, vol. 104, pp. 834 – 844, 2013.

[123] J. Kuo, D. Romero, and C. Amon, “A novel wake interaction model for wind farm layout opti-

mization,” in Proceedings of the ASME 2014 International Mechanical Engineering Congress &

Exposition. Montreal, Canada., pp. 1–10, 2014.

[124] J. Y. Kuo, D. A. Romero, and C. H. Amon, “A mechanistic semi-empirical wake interaction model

for wind farm layout optimization,” Energy, vol. 93, pp. 2157–2165, 2015.

Page 90: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

BIBLIOGRAPHY 80

[125] J. Feng and W. Z. Shen, “Solving the wind farm layout optimization problem using random search

algorithm,” Renewable Energy, vol. 78, pp. 182–192, 2015.

[126] Y. Chen, H. Li, K. Jin, and Q. Song, “Wind farm layout optimization using genetic algorithm

with different hub height wind turbines,” Energy Conversion and Management, vol. 70, pp. 56 –

65, 2013.

[127] Y. Chen, H. Li, B. He, P. Wang, and K. Jin, “Multi-objective genetic algorithm based innovative

wind farm layout optimization method,” Energy Conversion and Management, vol. 105, pp. 1318–

1327, 2015.

[128] A. Chehouri, R. Younes, A. Ilinca, and J. Perron, “Review of performance optimization techniques

applied to wind turbines,” Applied Energy, vol. 142, pp. 361 – 388, 2015.

[129] B. P. Rasuo and A. C. Bengin, “Optimization of wind farm layout,” FME Transactions, vol. 38,

no. 3, pp. 107–114, 2010.

[130] J.-O. Mo, A. Choudhry, M. Arjomandi, and Y.-H. Lee, “Large eddy simulation of the wind tur-

bine wake characteristics in the numerical wind tunnel model,” Journal of Wind Engineering and

Industrial Aerodynamics, vol. 112, pp. 11–24, 2013.

[131] Y.-T. Wu and F. Porte-Agel, “Large-eddy simulation of wind-turbine wakes: evaluation of turbine

parametrisations,” Boundary-layer meteorology, vol. 138, no. 3, pp. 345–366, 2011.

[132] H. Zhong, P. Du, F. Tang, and L. Wang, “Lagrangian dynamic large-eddy simulation of wind

turbine near wakes combined with an actuator line method,” Applied Energy, vol. 144, pp. 224–

233, 2015.

[133] E. S. Politis, J. Prospathopoulos, D. Cabezon, K. S. Hansen, P. Chaviaropoulos, and R. J.

Barthelmie, “Modeling wake effects in large wind farms in complex terrain: the problem, the

methods and the issues,” Wind Energy, vol. 15, no. 1, pp. 161–182, 2012.

[134] R. Mikkelsen, Actuator disc methods applied to wind turbines. PhD thesis, Technical University of

Denmark, 2003.

[135] F. Castellani and A. Vignaroli, “An application of the actuator disc model for wind turbine wakes

calculations,” Applied Energy, vol. 101, pp. 432 – 440, 2013. Sustainable Development of Energy,

Water and Environment Systems.

Page 91: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

BIBLIOGRAPHY 81

[136] N. Troldborg, J. N. Sørensen, and R. Mikkelsen, “Actuator line simulation of wake of wind turbine

operating in turbulent inflow,” in Journal of Physics: Conference Series, vol. 75, p. 012063, IOP

Publishing, 2007.

[137] L. A. Martınez, S. Leonardi, M. J. Churchfield, and P. J. Moriarty, “A comparison of actuator disk

and actuator line wind turbine models and best practices for their use,” AIAA Paper, no. 2012-

0900, 2012.

[138] F. Porte-Agel, H. Lu, and Y.-T. Wu, “Interaction between large wind farms and the atmospheric

boundary layer,” Procedia IUTAM, vol. 10, pp. 307–318, 2014.

[139] P.-E. M. Rethore, N. N. Sørensen, and F. Zahle, “Validation of an actuator disc model,” 2010.

[140] B. Blocken, T. Stathopoulos, and J. Carmeliet, “CFD simulation of the atmospheric boundary

layer: wall function problems,” Atmospheric environment, vol. 41, no. 2, pp. 238–252, 2007.

[141] J. Prospathopoulos, E. Politis, and P. Chaviaropoulos, “Modelling wind turbine wakes in complex

terrain,” Proceedings EWEC 2008, Brussels, Belgium, 2008.

[142] L. Davis, “Genetic algorithms and simulated annealing,” 1987.

[143] J. Annoni, P. Seiler, K. Johnson, P. Fleming, and P. Gebraad, “Evaluating wake models for wind

farm control,” in American Control Conference (ACC), 2014, pp. 2517–2523, IEEE, 2014.

[144] F. Castellani and A. Vignaroli, “An application of the actuator disc model for wind turbine wakes

calculations,” Applied Energy, vol. 101, pp. 432–440, 2013.

[145] G. Crasto, A. Gravdahl, F. Castellani, and E. Piccioni, “Wake modeling with the actuator disc

concept,” Energy Procedia, vol. 24, pp. 385–392, 2012.

[146] J. F. Ainslie, “Calculating the flowfield in the wake of wind turbines,” Journal of Wind Engineering

and Industrial Aerodynamics, vol. 27, no. 1, pp. 213–224, 1988.

[147] F. Castellani, A. Gravdahl, G. Crasto, E. Piccioni, and A. Vignaroli, “A practical approach in the

CFD simulation of off-shore wind farms through the actuator disc technique,” Energy Procedia,

vol. 35, pp. 274–284, 2013.

[148] J. Johansen and N. N. Sørensen, “Aerofoil characteristics from 3D CFD rotor computations,”

Wind Energy, vol. 7, no. 4, pp. 283–294, 2004.

[149] K. Rados, J. Prospathopoulos, N. C. Stefanatos, E. Politis, P. Chaviaropoulos, and A. Zervos,

“CFD modeling issues of wind turbine wakes under stable atmospheric conditions,” in European

wind energy conference & exhibition proceedings, Parc Chanot, Marseille, France, pp. 16–19, 2009.

Page 92: Enhanced Layout Optimization and Wind Aerodynamic Models ... · Enhanced Layout Optimization and Wind Aerodynamic Models for Wind Farm Design Yen Jim Kuo Doctor of Philosophy Graduate

BIBLIOGRAPHY 82

[150] W. L. Steffen and O. T. Denmead, Flow and Transport in the Natural Environment: Advances and

Applications: Advances and Applications. Springer Science & Business Media, 2012.

[151] P. Y. Zhang, J. Y. Kuo, D. A. Romero, T. C. Chan, and C. H. Amon, Advances and Trends in

Optimization with Engineering Applications, ch. Robust Wind Farm Layout Optimization. Society

for Industrial and Applied Mathematics.

[152] F. Porte-Agel, Y.-T. Wu, and C.-H. Chen, “A numerical study of the effects of wind direction on

turbine wakes and power losses in a large wind farm,” Energies, vol. 6, no. 10, pp. 5297–5313,

2013.

[153] P. Fleming, P. M. Gebraad, S. Lee, J.-W. Wingerden, K. Johnson, M. Churchfield, J. Michalakes,

P. Spalart, and P. Moriarty, “Simulation comparison of wake mitigation control strategies for a

two-turbine case,” Wind Energy, vol. 18, no. 12, pp. 2135–2143, 2015.

[154] A. Foley, P. Leahy, K. Li, E. McKeogh, and A. Morrison, “A long-term analysis of pumped hydro

storage to firm wind power,” Applied Energy, vol. 137, pp. 638–648, 2015.

[155] N. Li, M. Hedayati, and K. W. Hedman, “Utilizing flywheels to provide regulation services for

systems with renewable resources,” in Power & Energy Society General Meeting, 2015 IEEE,

pp. 1–5, IEEE, 2015.

[156] B. V. Mathiesen, H. Lund, D. Connolly, H. Wenzel, P. A. Østergaard, B. Moller, S. Nielsen,

I. Ridjan, P. Karnøe, K. Sperling, et al., “Smart Energy Systems for coherent 100% renewable

energy and transport solutions,” Applied Energy, vol. 145, pp. 139–154, 2015.