Вы находитесь на странице: 1из 93

Airfoil Optimization

for vertical axis wind turbines

Delft University of Technology

Rody Kemp

March 13, 2015

A IRFOIL O PTIMIZATION
FOR VERTICAL AXIS WIND TURBINES
by

Rody Kemp

in partial fulfillment of the requirements for the degree of

Master of Science
in Aerospace Engineering
Wind Energy Research Group, Faculty of Aerospace Engineering, Delft University of Technology

at the Delft University of Technology,


to be defended publicly on Wednesday March 26, 2015 at 10:30.

Thesis committee:

Dr. ir. C. J. Simao Ferreira (supervisor)


Ir. W. A. Timmer,
Ir. L. M. M. Boermans,

TU Delft Wind Energy


TU Delft Wind Energy
TU Delft Aerodynamics

An electronic version of this thesis is available at http://repository.tudelft.nl/.

A CKNOWLEDGMENTS
With this thesis an end has come to my years as a student. Obviously, I could not have done this without the
help of a lot of people. First of all I would like to thank my supervisor Carlos Simao Ferreira, he inspired me to
pursue VAWT research during my DSE, supported me during my internship and finally supervised my thesis.
Next to this I have had invaluable help with understanding the optimizer and airfoils in general from Gael
Andrade de Oliveira. Our (long) talks about airfoils, optimizers (and e-learning) made of lot of this thesis to
what it is now.
On a personal level I would first like to thank my girlfriend Judit, she has always been there for me. She even
read this thesis without understanding half of it, if that isnt love. Next thanks to my family; Martin, Marijke
and Cintha, for being proud on me and supporting me with big decisions. To my graduation buddy Dieter:
Thanks for distracting me from my work and making me drink more beer than is good for me.
If I am going to thank everyone that has been important to me my thesis will probably exceed the page limit
so I will not do that. Instead I will thank you all together for making my life here in Delft awesome. It was a
great ride with all of you with lots of very good memories.
Rody Kemp
Delft, March 2015

iii

S UMMARY
This thesis addresses the process of airfoil optimization for vertical axis wind turbines (VAWT). The airfoils
are designed for large scale turbines above 5 MW. The VAWT concept is relevant for offshore floating wind
energy, because of its low center of gravity (stability) and their simplicity (low maintenance). An optimal tip
speed ratio of 4-4.5 is chosen with an average Reynolds number of 5 million. The solidity c/R of the turbine is
0.1. These operation conditions are representative for the new generation VAWT. The goal of this thesis is to
develop an optimization process for VAWT airfoils and to demonstrate it by designing an airfoil, while taking
into account airfoil soiling.
A literature review presents the previous research in VAWT airfoil design, showing that no consensus has been
previously reached about VAWT airfoil design. From the literature review an optimization objective derived
by Simo Ferreira [30] is chosen. The airfoil is optimized for aerodynamic and structural performance. The
C

aerodynamics is assessed on airfoil level according to the objective of lift slope over drag ( Cl ). Structurally,
the airfoil will be optimized for flapwise bending stiffness
using turbulent transition.

I xx
ts .

Airfoil soiling is simulated on the airfoil by

A genetic optimization tool for airfoils coupled with RFOIL, an airfoil analysis tool, is used to generate VAWT
airfoils. The objective function values are calculated using the aerodynamic coefficients from RFOIL and
the geometric properties of the airfoils. The optimization process is validated by analyzing the results with
three different models for full VAWT analysis. These models are: 1) an inviscid panel model coupled with
RFOIL, 2) a double wake panel model and 3) a CFD model. Three airfoils resulting from the optimization
are tested using the aerodynamic models. The performance of the airfoils validates the objective functions,
but performance for the soiled case is not satisfactory. These preliminary findings were presented at the 33rd
Wind Energy Symposium at the AIAA SciTech conference [32], the full paper can be found in appendix B
Five different strategies are developed to optimize airfoils. The results are analyzed using a double wake
panel model. The optimization strategy in which airfoils are optimized for soiled conditions results in the
best performing airfoils. The RK2-27 is a demonstration airfoil resulting from this optimization strategy. The
C P of this airfoil for a tip speed ratio of 4 and a solidity of 0.1 is 0.53 in the clean case and 0.45 in the soiled
case. This was determined by both the inviscid panel model and the double wake panel model.
The RK2-27 has an increased C P compared to the NACA 0018 of 0.04 in the clean case at the design operating
conditions. The C P in the soiled case is only 0.02 lower than the NACA 0018. The maximum thickness of the
airfoil increased by 50% from 18% to 27%. The RK2-27 has similar aerodynamic performance compared to
the traditionally used NACA 0018, while structurally it performs significantly better.

C ONTENTS
Acknowledgments

iii

Summary
1 Introduction
1.1 Wind Energy . . . . . . . .
1.2 Vertical Axis Wind Turbines .
1.3 Research Goal . . . . . . . .
1.4 Overview of Thesis . . . . .

v
.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

2 Literature Review
2.1 Historical overview: 1970-1990 . . . . . . . .
2.1.1 Wichita State University . . . . . . . .
2.1.2 Sandia National Laboratories . . . . .
2.1.3 Queens University of Belfast . . . . . .
2.1.4 Tokai University . . . . . . . . . . . .
2.1.5 West Virginia University . . . . . . . .
2.1.6 National Technical University of Athens
2.1.7 University of Glasgow . . . . . . . . .
2.1.8 Griffith University . . . . . . . . . . .
2.2 Recent developments: 2005-present . . . . .
2.2.1 Delft University of Technology . . . . .
2.2.2 University of Windsor . . . . . . . . .
2.2.3 Numerical Methods . . . . . . . . . .
2.2.4 Commercial developments . . . . . .
2.3 Conclusion . . . . . . . . . . . . . . . . . .
3 Airfoil Design Methodology
3.1 Airfoil optimization goals . . .
3.1.1 Aerodynamic Function.
3.1.2 Structural Function . .
3.2 Roughness modeling . . . . .
3.3 Optimization strategy . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

1
1
1
3
4

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

5
7
7
7
8
9
9
10
11
11
12
12
13
15
16
16

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

19
19
19
21
21
22

4 Airfoil Genetic Optimization


4.1 Optimizer . . . . . . . . . . . . . . . .
4.1.1 Parameterization. . . . . . . . .
4.1.2 Matlabs Genetic Optimizer . . .
4.1.3 RFOIL . . . . . . . . . . . . . .
4.2 Constraints . . . . . . . . . . . . . . .
4.3 Implementation of objective functions .
4.3.1 Aerodynamic objective function .
4.3.2 Structural objective function . . .
4.4 Discussion on Inputs and Constraints .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

25
25
25
27
29
30
31
31
32
33

5 Airfoil Performance Analysis


5.1 Inviscid Panel Model . . . . .
5.2 Double Wake Panel Model . .
5.3 CFD Simulation . . . . . . . .
5.4 Comparison between models .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

35
35
37
37
38

.
.
.
.
.

.
.
.
.

.
.
.
.
.

.
.
.
.

.
.
.
.
.

.
.
.
.

.
.
.
.
.

.
.
.
.

.
.
.
.

vii

viii
6 Results of Optimization
6.1 Optimization Strategy 1 . . . . .
6.1.1 Inputs optimization . . .
6.1.2 RK1-23 . . . . . . . . . .
6.1.3 RK1-26 . . . . . . . . . .
6.1.4 RK1-32 . . . . . . . . . .
6.1.5 Discussion of strategy 1 .
6.2 Adapted optimization strategies

C ONTENTS

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

41
41
41
42
42
43
44
47

7 Final Remarks
53
7.1 Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
7.2 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
Bibliography

55

A Coordinates of Results

59

B AIAA paper

65

N OMENCLATURE
Abbreviations
AoA

Angle of attack

HAWT

Horizontal Axis Wind Turbine



Tip speed ratio VR
1

TSR
VAWT

Vertical Axis Wind Turbine

Greek Symbols

Angle of attack of airfoil

Bound circulation on airfoil

Inflow angle at airfoil

Density (of air)

Azimuthal angle of VAWT

t ai l

Trailing edge angle

Latin Symbols
Bn

Bernstein polynomial coefficient

Cd

Drag coefficient of airfoil

Cl

Lift coefficient of airfoil

@C
Lift slope of airfoil @l

Cl
I xx

Moment of inertia around neutral axis of airfoil

Ncr i t

Amplification factor for XFOIL/RFOILs transition model

sc aer o

Final aerodynamic objective function value

sc cl ean

Aerodynamic objective function value in clean case (free transition)

sc r oug h

Aerodynamic objective function value in rough case (forced transition)

TD

Torque due to drag force

ts

Skin thickness of airfoil

t ai r (x)

Airfoil relative thickness

U1

Incoming (wind) speed

Vp

Perceived velocity by the airfoil

wc

Weighting factor of clean case.

ws

Weighting factor of soiled case.

xt r

Chordwise location where forced transition is applied in XFOIL/RFOIL

zt e

Trailing edge thickness

Number of blades of turbine

Chord of airfoil

order of Bernstein polynomial

Normalized x coordinate on x = [0,1]


ix

1
I NTRODUCTION
In this thesis, an approach to airfoil optimization for vertical axis wind turbines (VAWT) is outlined. First, a
short introduction into VAWT and VAWT aerodynamics will be given. Also an overview of the research goals
and the ouline of this thesis will be presented.

1.1. W IND E NERGY


The world is facing a great energy challenge. With increasing population and increasing wealth, the demand
for energy is increasing fast. The diminishing supplies of fossil fuels and the threat of global warming call for
sustainable solutions. These solutions are found in a new sustainable energy mix consisting of hydro-, solarand windpower. Wind energy, especially in Europe, South-east Asia and North America, is a fast growing
industry and will play a significant role in the solution of the energy challenge the world is facing.
Space and environmental constraints force the wind energy generation to move offshore. Onshore projects
are still developed, but are often confronted with resistance from people living nearby complaining about the
noise and shadow nuisance. In highly populated areas simply no space is available to place enough turbines
to facilitate energy for everyone. Offshore, the wind farms can be made bigger and are not hindered by noise
and shadow constraints. The drawbacks of offshore wind are higher installation costs and higher maintenance cost because the sites are harder to reach. Next to this, the foundation of turbines can be a problem.
In shallow waters the turbines can be fixed to the ground, but for deep waters this becomes too costly. In
offshore environments, VAWT have some beneficial characteristics which make them an interesting research
subject.

1.2. V ERTICAL A XIS W IND T URBINES


In the field of wind energy, the most well-known and widely used turbines are the horizontal axis wind turbines (HAWT) . This turbine is being used onshore as well as offshore and is available in all sorts of sizes
ranging from 1 kW up to 10 MW, with plans of even larger scales. VAWT are not well-known and not widely
used yet, only small turbines in the order of 10 kW are commercially available and bigger turbines are still in
development. But, this has not always been the case and might not be the case anymore for very long.

H ISTORY
VAWT have been present for a long time. As long as 1500 years ago a primitive version was used by the
Persians, called a Panemone. These windmills were used to grind crops or pump water. The Panemone
used sails or wooden surfaces to turn around a vertical axis. The windward turning part of the windmill was
shielded by a wall to avoid it from producing drag. An example of a Panemone can be found in figure 1.1a.
The first modern VAWT for electricity generation were made during the 1970s and 80s. During this time, wind
energy was in its starting days and different concepts were competing to become the preferred choice. Up

(a) Panemone using wood as sails [1].

1. I NTRODUCTION

(b) Schematic of Savonius rotor [2].

(c) Schematic of Darrieus rotor [3].

Figure 1.1: Different types of VAWT

until now the most successful VAWT in history was produced by a spin-off of the Sandia National Research
Laboratories in the US. They had a fleet of about 500 vertical axis turbines in service. These FloWind turbines
were troposkien () shaped turbines with a rated power of about 150-250 kW. More on this turbine can be
found in section 2.1.2 and a picture of this turbine is printed on the cover of this thesis.
Around the start of the 90s, research was mainly focused on the development of the HAWT. This led to the
shutdown of most VAWT research programs while HAWT programs were boosted. This is why VAWT nowadays still lack behind their horizontal counterparts on which a few decades more of research has been devoted. In theory, both concepts should be able to perform similarly.

VAWT CONCEPTS & WORKING PRINCIPLES


There are a variety of VAWT concepts, a first distinction can be made between drag driven and lift driven
turbines. Drag driven turbines use the aerodynamic drag force to generate energy. The most famous example
of this principle is given by a simple cup wind meter (anemometer) which uses the different in drag coefficient
of a convex and concave sphere to turn. A schematic of a drag driven Savonius turbine is shown in figure 1.1b,
in this case the convex and concave drag characteristics are also used. Since these machines can attain only
very low efficiency they are mainly used for high torque applications. They will not be considered in this
thesis.
The focus of this thesis will be on lift driven VAWT, which will be referred to as just VAWT for simplicity.
Theoretical analysis shows that these turbines can attain the same efficiency as HAWT. These VAWT are driven
by the aerodynamic lift force acting on the blades. These blades have airfoil shapes and a schematic of such a
rotor can be seen in figure 1.1c. There are a lot of shape variations possible when it comes to these VAWT. The
most simple one is the H-shaped (figure 1.1c), this concept can also be twisted into a helical shape. Another
promising one is the shaped turbine, the FloWind turbine from the 80s had such a shape. A less widely
used concept is the V-shaped VAWT. In this thesis a 2D approach is taken, which implies that it can still be
used in all turbines.
Since the airfoils of a VAWT are constantly changing their orientation with respect to the incoming wind, they
experience changing wind speeds and angles of attack. This makes the operating regime of VAWT unsteady
and the variation in angle of attack means that an airfoil should perform well at a big range of angles instead
of a small range like with HAWT. A VAWT is driven by torque that is generated by the blades, this torque is
produced by the lift force on the airfoil, a schematic of velocities and forces is shown in figure 1.2. Airfoils
produce different lift at different angles of attack. Dynamic effects like dynamic stall make the aerodynamics even more complex. On top of this, a VAWT blade passes through its own wake, created on the upwind
part of the rotation, during the downwind part of the rotation. This blade-wake interaction makes the power
extraction downwind very different than upwind. The goal of this thesis is to design an airfoil which is aerodynamically efficient and is structurally strong enough to handle the forces on the blade.

1.3. R ESEARCH G OAL

Figure 1.2: The working principles of a lift driven VAWT [4].

R ELEVANCE OF VAWT RESEARCH


After this introduction to VAWT, the question could rise: Why investigate VAWT when HAWT are already far
more advanced and working fine? Before this question is answered the assumption is made that wind energy
will have to grow fast in the coming decades to be able to provide enough green energy to substitute fossil
fuels. In this growth process, VAWT could have some advantages over HAWT.
An area where VAWT are already applied is small scale power generation. Since VAWT can capture wind
from all directions, urban areas are also suitable for VAWT. Commercial products are already available
in this range [5].
The most promising area for VAWT application however, is offshore wind energy. Currently wind turbines are placed on the seabed on big submerged structures, which are only worthwhile in relatively
shallow waters. Floating turbines are a huge opportunity to not only make installment easier, but also
to expand the area where turbines can be placed. When looking at floating turbines, VAWT have advantages over HAWT like a lower center of gravity and load application point, which would make the
floater much more stable.
Another advantage of a VAWT is the fact that upscaling of the technology is easier. Right now the blades
of HAWT are getting to their limits in terms of length and twist. VAWT blades can be supported by more
than one strut and are thus less limited in size. It has to be noted here that a VAWT is usually more
expensive in material when you compare it to a same rotor area HAWT. This is because VAWT rotors are
three dimensional and hence mass scales with radius to the third power, while for HAWT it is a 2D rotor
and thus it scales with the radius to the second power.
VAWT show a good performance in skewed flow. This is because their actual frontal area increases. This
is a positive feature when installing turbines on rooftops but more importantly when installing VAWT
on floating platforms subject to waves.
These considerations make the VAWT a viable candidate to expand the possibilities in wind energy and hence
a contribution to solve the challenge of diminishing fossil fuel supplies.

1.3. R ESEARCH G OAL


The overall goal of this thesis is to lower the cost of energy produced by a VAWT by focusing on the aerodynamic efficiency of the airfoils and the structural possibilities of the airfoils. The specific goal of this research
is twofold.
1. To design an airfoil shape or family of shapes that is suitable for VAWT. This will give a general idea on
what VAWT airfoils should look like and what characteristics are advised.

1. I NTRODUCTION
2. To validate the objective functions and optimization algorithm as an efficient optimization method to
design airfoils in future projects. For instance for different Reynolds numbers or different operating
conditions of the turbine.

The turbine considered in this research will be a large scale VAWT. As a result, the Reynolds number is taken to
be 5 million. The aerodynamics of a VAWT are such that the Reynolds number is not constant, but 5 million is
taken as a representative mean value. Higher Reynolds numbers lead to higher stall angles which is beneficial
for VAWT due to their wide angle of attack operation range. The solidity that is used as a reference is 0.1, for
a two bladed turbine the solidity is computed using the chord to radius ratio c/R. The solidity should be
similar between turbines to compare performance. The tip speed ratio of the turbine is taken to be 4-4.5.
This is where theoretically the highest efficiency can be reached. Figure 1.3 shows this.

Figure 1.3: The maximum efficiency of a VAWT is reached at a tip speed ratio of 4-4.5 [6].

1.4. O VERVIEW OF T HESIS


In this thesis first a literature review of previous VAWT airfoil research will be given in chapter 2. Then the
theoretical foundation for the optimization will be discussed in chapter 3. The optimization tool with all its
inputs will be presented in chapter 4. The models with which the results of the optimization are be validated
can be found in chapter 5, together with a comparison between all the models. Finally the results will be
analyzed in chapter 6. The thesis will end with a conclusion of the research in chapter 7. This chapter will
show the main findings of the research and give recommendations on how to use the optimization. Also
recommendations for further research will be given in this chapter.

2
L ITERATURE R EVIEW
Wind energy is moving offshore. Floating wind turbines is the newest development in the field and this has
sparked new interest in vertical axis wind turbines (VAWT) because of their beneficial characteristics for floating structures.
The VAWT concept should theoretically be able to reach similar aerodynamic performance as the horizontal
axis wind turbine (HAWT). Compared to HAWT, an apparent advantage of VAWT is that it is less complex and
should therefore be less costly in maintenance, which is a big part of the cost for offshore wind. Next to this
the VAWT concept seems more compatible to floating platforms because of its lower center of gravity, which
will increase stability and decrease the required size of the floater.
Much research has been done on VAWT in the 1970s and 1980s, after this the HAWT became the dominant
concept and research efforts for VAWT faded. Recently the interest has returned. In this chapter, an attempt
is made to summarize all previous research regarding airfoils for VAWT. For VAWT the airfoil can have a huge
impact in the cost of energy (COE) of the turbine. Both aerodynamic efficiency and structural integrity of the
blades rely on the choice of airfoil. The unsteady operation of the VAWT, with a constant variation in apparent
speed and angle of attack seen by the airfoil, complicate the design of the airfoil.
This chapter will present the research done in airfoil design for VAWT. This chapter will provide an overview
of clear rationales behind airfoil design and might help to provide new insights into existing challenges. The
chapter will show the main objectives of the research, the method of design and analysis and the final results
and potential implementation.

Figure 2.1: The amount of papers published on VAWT research (based on data retrieved from [7]).

The review is divided in two parts. Figure 2.1 shows that there are two distinct time frames where most research took place on VAWT. These two time frames are separated in this work. Table 2.1 shows the VAWT
airfoil research done in chronological order.

2. L ITERATURE R EVIEW
Table 2.1: List of research done on VAWT airfoils.

Year

Author

Airfoil

Main objectives or subject

1977
1978
1978

T. Fukuda
E.G. Kadlec [9]
J.V. Healy [10, 11]

WSU 0015
NACA 00xx
NACA 00xx, G
profiles

Not found, experimental work on airfoil used [8].


First Sandia report on airfoils for VAWT.
The influence of thickness and camber on VAWT performance.

1980

Y. Kato, K. Seki, Y.
Shimizu [12]

TWT 11215-1

1981

R.E. Sheldahl, P.C.


Klimas [13]
P.G. Migliore, J.R.
Fritschen [14]

NACA 00xx

Airfoil design objectives: 1) High @l , 2) Low C d , 3)


C d should be symmetric about C l =0 , 4) Large negative
pitching moment.
Full 360 wind tunnel tests of NACA 00xx

1983

1984

P.C. Klimas [15]

1988

A. Zervos [16]

1990

D.E. Berg [17]

1992

T.D. Ashwill [18]

1992

R.A.M.
Galbraith,
F.N.
Coton,
J.
Dachun [19]
B.K. Kirke [20]
J.C. Vassberg, A.K.
Gopinath, A. Jameson [21]
M.C. Claessens [22]

1998
2005

2006

2007

2007

2010
2011
2012

R. Bourguet,
G.
Martinat, G. Harran,
M.Braza [23]
M. Islam, D. Ting, A.
Fartaj [24]

T.J. Carrigan [25]


M.R. Castelli, E.
Benini [26]
H.J. Sutherland [27]

NACA
00xx,
NACA 6-series,
WSU 0015
SAND 0015/47,
SAND 0018/50,
SAND 0021/50
ARC 0015, NACA
00xx, NACA 6series, GAW(1),
NLF0416
SAND 0015/47,
SAND 0018/50,
SAND 0021/50
NACA
0021,
SAND 0018/50
GUVA 10

NACA
0015,
WARP0015-RC8
DU06W200

unnamed

NACA
0015,
GOE 420, NACA
4415,
NASA
LS-0417, NASA
NLF-0416,
S1210,
MIVAWT1
unnamed
NACA
0012,
NACA 0021
SNLA
series,
S824

@C

Review of existing airfoils with desirable characteristics:


@C
low C d and high @l .
Favourable characteristics: 1) modest values of maximum lift coefficient with sharp stall, 2) low C D 0 , 3) wide
drag buckets.
Review of existing airfoils stressing flow curvature effects.

Discussion of SAND airfoils and robustness to roughness.


The 34 meter testbed turbine of Sandia.
Airfoil designed for stall regulation.

Selfstarting VAWT.
Warped airfoil design on basis of CFD.

Airfoil design criteria: 1) Low Reynolds (105 ), 2) Low C d ,


3) wide drag bucket, 4) Increased thickness for structural strength, 5) smooth stall for noise reduction, 6)
small hysteresis loop, 7) postpone deep stall and small
drop in C l .
Optimization using commercially available tools.

Review of existing airfoils and design of new airfoil.

Genetic optimization using NACA 4-series.


A comparison using CFD between two airfoils to investigate the effect of relative thickness.
A review of Sandias VAWT research.

2.1. H ISTORICAL OVERVIEW: 1970-1990

2012

L.A. Danao, N. Qin,


R. Howell [28]

2012

T.J. Carrigan, B.H.


Dennis, Z.X. Han,
B.P. Wang [29]
C.J. Simao Ferreira,
B. Geurts [30]
C.J. Simao Ferreira
et al. [32]

2014
2015

NACA
NACA
NACA
LS0421
NACA
NACAopt

0012,
0021,
5522,

The effect of thickness and camber is investigated using


the NACA 4 series.

0015,

Based on his previous work ([25]).

DU12W262
unnamed

Genetic optimization [31] airfoil design with objective


lift slope over drag.
Follow up on previous research.

2.1. H ISTORICAL OVERVIEW: 1970-1990


In the 1970s and 80s wind energy was emerging as a consequence of the energy crisis. The concepts for
VAWT and HAWT were competing to become the industry standard and a lot of research was done into both
concepts. This period is when the first big research efforts were done into (lift driven) VAWT technology.
In this section, these efforts will be reviewed. They are categorized by institution and will be presented in
chronological order.

2.1.1. W ICHITA S TATE U NIVERSITY


In 1977 a master thesis project on Wichita State University focused on a new airfoil design for Darrieus wind
turbines. The thesis itself is not publicly available but experimental validation of this work was done in the
WSU wind tunnel using a one meter diameter, three bladed turbine with troposkien shaped blades [8]. The
main objective for the design of the airfoil of the blade found by Fukuda [33] was increasing the stall angle of
the airfoil. It was found that this is mainly done by increasing the thickness of the airfoil. This also makes the
peak in performance shift towards lower TSR.
The airfoil that was designed in the thesis was the WSU 0015 and the WSU 0021, which are symmetric profiles.
The WSU 0021 is shown in figure 2.2. The experimental comparison with the NACA 0012 gave the following
results, the NACA 0012 had a C P,max of about 0.2 at a TSR of 5 and the WSU 0021 achieved a C P,max of about
0.25 at a TSR of 4.

Figure 2.2: The WSU 0021 airfoil [8]

2.1.2. S ANDIA N ATIONAL L ABORATORIES


The main research institute during the 70s and 80s was Sandia National Laboratories (SNL) based in New
Mexico, USA. The research on VAWT was commissioned by the US Department of Energy and a lot of resources were available to carry out research. At SNL, the popularity of symmetric NACA 4-digit airfoils originates. The early research focused on the NACA 0012 and NACA 0015 [9, 13]. These airfoils were originally
designed for aviation purposes and were not specifically designed for VAWT. A symmetric airfoil was chosen because of the operational nature of the VAWT, the suction and pressure side of the airfoil change sides
during the revolution. It was deemed important that the behavior of the airfoil sections remained the same,
so symmetric sections were chosen. The NACA 00xx series was chosen mainly due to the amount of data
available on these airfoils. A complete dataset for the NACA 00xx series was published [13] after an extensive

2. L ITERATURE R EVIEW

measurement campaign in which the airfoils were tested over the full range of angles of attack from -180 to
180 .
When SNL started to work on larger prototypes of 17 meter (and later 34 meter) the need for more structural
stiff airfoils arose. For this reason the researchers switched to the NACA 0018, which has a higher thickness. At
the same time new airfoil shapes were investigated by Klimas [15]. Klimas stated the desirable characteristics
of a good VAWT airfoil:
Modest values of maximum lift coefficient and sharp stall.
Low C D,0
Wide drag buckets
The first point is desirable for power regulation of the turbine at high windspeeds, the modest value of C l
reduces peak loads in these conditions and the sharp stall will aid in slowing down the turbine. The second
point and the related third point originate from the unsteady operation of the VAWT where the angle of attack
varies. Low and wide drag buckets ensure efficient operation over large ranges of angle of attack.
Airfoils possessing the characteristics mentioned above are NLF airfoils. In light of these findings SNL designed three NLF airfoils in cooperation with Gregorek of Ohio State University [17]. Use was made of the
airfoil design code PROFILE by Eppler [34]. The airfoils were based on the NACA 00xx series and are called
the SNLA (or SAND) 0015/47, 0018/50 and 0021/50. The first four digits indicate the symmetry and thickness
and the last two digits the location in percentage of chord two where the flow is laminar. Advantages of the
SNLA family over the NACA 00xx family include reduced roughness sensitivity and thus higher performance
with dirty blades.
SNLs final 500 kW Test Bed turbine was a troposkien shaped VAWT with a maximum diameter of 34 meters.
This turbine utilized a NACA 0021 near the towers for structural strength and the SNLA 0018/50 near the
equatorial region for aerodynamic performance [27].
The Test Bed turbine was commercialized by FloWind Corporation, using a SNLA 0021/50 instead of a SNLA
0018/50 profile. During the peak of FloWind, 500 turbines were in service. A new operating rpm led to a need
of a slightly modified profile. This profile was designed by Dan Somers [35] and was called the S824 [36]. This
profile was used in the B blade and C blade of the FloWind turbines [37].

2.1.3. QUEEN S U NIVERSITY OF B ELFAST


At Queens University of Belfast, Healy [10, 11] conducted an investigation on the effect of airfoil thickness and
camber on the power output. The investigation was done for relatively small turbines with solidities from 0.1
to 0.2 and Reynolds numbers of about 4 105 . The main objective of this research was to investigate the effect
of airfoil thickness and camber for the purpose of curved blade VAWT. In this research, no airfoil was designed,
but several airfoils were tested using a single actuator multiple streamtube model. Healy acknowledges the
fact that this model does not have a correct treatment of the downwind halve of the turbine but states that
satisfactory double multiple streamtube models have not been devised successfully.
The airfoils that were used for the investigation into airfoil thickness effects were the NACA 0009, 0012, 0015
and 0018 [11]. Airfoil coefficient data was taken from the NACA TR-586 report [38]. The main result of this
work is that airfoil thickness does not play a large role in the maximum power output at high Reynolds numbers in the order of 2 million. For lower Reynolds, thicker profiles are preferred.
The second investigation into camber took into account Gttingen series airfoils, for the reason that for these
airfoils data was available to do the calculations. This data was taken from [39]. Profiles were chosen to
have different cambers and were compared to two symmetric profiles, the NACA 0018 and the G 460. The
cambered profiles were G 676, 738, 735, 746, 741, 744 and 420. Of these airfoils, only the G 735 and 738
are considered interesting because higher power coefficients were achieved and the C p T SR curves were
smooth, making it easier for the turbine to operate in optimum conditions. It was found that both symmetric
profiles had the best performance in terms of smoothness of power curve and height of C P,max . It was concluded that the closer the airfoil is to symmetric, the more satisfactory the power output. All airfoils were also
tested at different pitch angles. It was found that most profiles have a big sensitivity to this preset pitch angle.

2.1. H ISTORICAL OVERVIEW: 1970-1990

2.1.4. T OKAI U NIVERSITY


In light of the 1970s energy crisis, countries where natural resources were limited invested in research on wind
energy. In 1980 Tokai University in Japan carried out research on airfoils for VAWT. The analysis and experimental work was performed on a small 22.5 meter (HD) turbine. A set of conditions for high aerodynamic
efficiency was derived [12]:
Large lift gradient

dC L
d

C D must be small and symmetric about C L =0


Large pitching moment coefficient
These objectives were derived using a simple streamtube model of the turbine. The dependency of C P on the
airfoil coefficients was analyzed. C L , C D and C M were considered. Having a large pitching moment coefficient is stressed in the paper as the most important design driver. An airfoil was designed with a S-shaped
camberline to introduce a strong negative pitching moment coefficient on the airfoil. The front part has a
negative camber while the aft part of the airfoil has a positive camber. This airfoil, the T.W.T.11215-1, was
eventually patented [40], but no records of commercial use were found.

Figure 2.3: Experimental results from Kato [12] showing the power coefficient (C P ) vs. tip speed ratio (). An increase in C P,max was
found for the airfoil with the pitching moment.

Experiments on the performance of the new airfoil were performed on the small experimental turbine. The
resulting C P -TSR curve is shown in figure 2.3. It is shown that the C P,max is increased when the airfoil has
a pitching moment. In the figure a positive C M is given, while the T.W.T.11215-1 has a negative pitching
moment. This is probably a notation error.

2.1.5. W EST V IRGINIA U NIVERSITY


Traditionally the NACA 00xx series were used for VAWT, these airfoils were designed for aeronautical applications. In the early 80s, an investigation was done at West Virginia University into other existing airfoils to
see if better performing airfoils were already available [14]. The performance was calculated using a blade
element momentum model by Strickland [41]. A selection of airfoils was made to limit computational efforts.
The selection

was based on the identified desirable airfoil characteristics, which were low drag and high lift

L
slope dC
d . Based on these requirements 10 airfoils were selected to be evaluated. The NACA 00xx, the NACA
63b 0xx, the NACA 64b 0xx and the WSU 0015 were chosen. The NACA series were all evaluated for the relative
thicknesses of 12%, 15% and 18%.

The calculations showed that the NACA 63 and 64 series had higher maximum C p over a higher TSR range.
For these series the aerodynamic performance increased with increasing thickness, which is favorable for
the structural stiffness of the blade. The NACA 632 015 and 633 018 performed the best of all airfoils, with the
NACA 632 015 reaching an annual energy output of about 20% more in comparison with the NACA 0015.
Migliore is also known for his work on flow curvature effects [42]. Using the methods outlined in [43] he trans-

10

2. L ITERATURE R EVIEW

Figure 2.4: Transformed shape of the NACA 632 015 to correct for flow curvature effects [14].

formed the NACA 632 015 to its appropriate shape. This was done for a c/R of 0.07. The original and resulting
profile can be seen in figure 2.4. It is interesting to note that this study concluded by favoring an airfoil that
was designed for maximizing laminar flow. This is in line with the findings and approach of Sandia.

2.1.6. N ATIONAL T ECHNICAL U NIVERSITY OF ATHENS


During the late 80s research on VAWT airfoil was performed at NTUA. The main focus of the airfoil design was
the influence of the rotational motion that the blade makes during its revolution and how this can be used
to balance the unsteady loading on a VAWT to increase the fatigue life. Cambered airfoils could help exploit
the rotational effect for this purpose. Other researchers have also suggested this influence and proposed
correction methods for VAWT analysis [14, 4244]. Zervos [16] suggested to use this rotational influence as a
basis for airfoil design for VAWT.
First a numerical evaluation was made on six airfoils to find the effect of thickness and camber on the unsteady loads in the VAWT. These six airfoils were: NACA 0012, NACA 0015, NACA 0018, NACA 63015, GAW(1)
and the NLF 0416. The effect of thickness on the unsteady loading was found to be insignificant. Camber did
have an effect on the behavior of the VAWT, Zervos explained this through the flow curvature effect, which
makes the same airfoil in a curved flow behave differently than in a rectilinear flow. A transformation to the
airfoil is made to correct for this effect by introducing a virtual angle of incidence and a virtual camber to the
airfoil, as can be seen in figure 2.5.

Figure 2.5: The virtual effect of flow curvature on the airfoil according to [16, 42]

Changing the (virtual) camber and incidence of a VAWT airfoil can shift loading from upwind to downwind
rotations. Zervos devised an airfoil that countered the flow curvature effect by cambering it in the opposite
way (concave inwards). The camberline of the airfoil was shaped according to the arc of the circle it was flying
on, the thickness distribution was that of a NACA 0015. This airfoil was called the ARC 0015 and since the flow
curvature effect changes with c/R, the airfoil is tailored to a specific c/R.

2.1. H ISTORICAL OVERVIEW: 1970-1990

11

2.1.7. U NIVERSITY OF G LASGOW


The wind energy community in Britain believed that large scale VAWT could be economically more favorable then the HAWT. Glasgow University worked on creating VAWT design codes which focused on the main
aerodynamic features of the VAWT. The code was capable of assessing the effect of blade pitch, twist, taper,
but also of airfoil section on the performance of the turbine. The assessment for the airfoils was done for a
straight bladed H-type turbine.
The main goal of the airfoil design was to increase the fatigue life of the turbine through passive stall regulation while retaining the behavior of the NACA 0018 during normal operation [19]. Stall regulation was to be
accomplished by having a sharper drop in C L at stall.
For the design of the new airfoil possessing these characteristics a design package was made. This package
contained two main components; an airfoil geometry generation module and a VAWT performance module.
Using this package and the GUVA 3 airfoil as a basis, a serie of six airfoils was created, the GUVA 4 to 9. Of
these six, the GUVA 4 and GUVA 9 had the most desirable static performance. During the assessment of the
airfoils, it became clear that the airfoils were both susceptible to dynamic stall and were likely to produce
some degree of stall regulation on a VAWT [19].
The GUVA 10 airfoil resulted from the design process. This airfoil was a combination of the NACA 0018 rear
part and the specially designed nose part of the GUVA 9. The rear part of NACA 0018 was used to achieve
more similar performance between the GUVA 10 and the NACA 0018. The final airfoil is presented in figure
2.6.

Figure 2.6: GUVA 10 airfoil (sharper nose) compared to the NACA 0018 airfoil [19].

The design process was validated with experiments on the GUVA 10 airfoil. Both static and dynamic tests
confirmed the static and dynamic stall behavior predicted in the design process. The GUVA 10 stalls approximately two degrees earlier than the NACA 0018. Combined with the sharper stall, this should provide stall
regulation on a VAWT.

2.1.8. G RIFFITH U NIVERSITY


A much debated subject in the wind energy community is the ability of a VAWT to self start. An interesting
approach to this discussion was taken by Kirke in his PhD thesis [20]. He investigated how a VAWT blade could
be designed to self start. Although he primarily focused on variable pitch mechanisms, some interesting notes
are made on airfoil design. A first interesting note is that for VAWT operating at high Reynolds numbers, the
self starting problem seems to be non existent. For small VAWT with very low Reynolds numbers a dead band
exists where negative torque is produced. The effect of Reynolds number can be seen in figure 2.7.

Figure 2.7: C P -TSR curve for different Reynolds numbers [20].

12

2. L ITERATURE R EVIEW

On airfoil shapes the following conclusions are drawn. First of all, thicker airfoils will decrease the size of
the dead band. The second conclusion is that cambered airfoils are favorable for self starting small wind
turbines. This is because performance of symmetric sections drops below Reynolds numbers of 2.5 105 , the
performance of cambered airfoils remains good. Another reason is that cambered airfoils will be able to extract more energy on the upwind pass because of higher lift coefficients there, since the incoming windspeed
here is higher, this would outweigh the loss in performance on the downwind pass.
The third and most exotic option opted are shape transforming blades. Three types are considered, fabric
sails, sheetmetal blades and flexible airfoils. The idea is that the camber reverses when the sign of the angle of attack changes. Fabric sails and flexible sheetmetal plates will have a natural tendency to do this, but
their basic aerodynamic performance lacks. At the time, flexible airfoils were deemed very interesting, with
self starting capabilities and 40% increase in peak performance, but too hard to manufacture. With recent
developments and advances in structural and material engineering, this option might be interesting to reinvestigate.

2.2. R ECENT DEVELOPMENTS : 2005- PRESENT


As can be seen in figure 2.1, around 2006 VAWT research started to grow rapidly. New application of wind
energy, in particular in floating offshore form, made VAWT an interesting concept again. Its advantages of
lower maintenance, low center of gravity and good performance in skewed flow are perfect for offshore floating conditions. The growth in computational power also changes the way in which designs are made, which
also marks a big difference between the two periods of VAWT research.

2.2.1. D ELFT U NIVERSITY OF T ECHNOLOGY


At TU Delft a lot of research is being done on VAWT at the moment. This section presents the published
results on airfoil design for VAWT until now.
C LAESSENS
The wind energy group at TU Delft has substantial resources dedicated to VAWT research. In 2006 a master
thesis was done on the optimization of an airfoil for a small scale VAWT [22]. The airfoil was designed for
the Turby VAWT, which is a small 2.652 meter (HD) VAWT for urban environments [45]. The goal for the
optimization was to design an airfoil with low drag and wide drag bucket. Contrary to the the early research
of Sandia and Glasgow University, the goal of smooth stall was set with the purpose of reducing noise. Also a
stiffness increase with respect to the NACA 0018 was desired.
Claessens conducted a simulation campaign in which several existing airfoils were tested in a double multiple streamtube (DMST) model to assess their performance. NACA 4 series, NACA 6 series, NLF airfoils and
profiles from the early VAWT period like the SNLA 0018 and the S824 were evaluated. Next to this the effect
of camber and thickness was evaluated using the NACA 4 series. It was concluded that NLF airfoils would be
the best basis for the design because of the low drag coefficients combined with relatively wide drag buckets.
With the NLF profiles as basis, first the thickness and camber are set. This is done by varying them and
analyzing them with the DMST model. Optimum thickness and camber are found to be 20% and 0.8% respectively. The pressure distribution and geometric features of the NLF profile are modified to improve the
drag characteristics of the airfoil. This resulted in the final DU06W200 profile which is shown in figure 2.8.

Figure 2.8: The final design of the DU06W200 compared with the NACA0018 [22]

Wind tunnel tests show that the DU06W200 behaves approximately the same as the NACA 0018 at negative
angles of attack. However at positive angles of attack the C L,max is increased. The overall drag bucket is wider

2.2. R ECENT DEVELOPMENTS : 2005- PRESENT

13

for the DU airfoil. This results in a C p curve shown in figure 2.9. An increase in C p,max is observed as well as
a slight shift towards the lower tip speed ratios.

Figure 2.9: The power coefficient (C p ) against the tip speed ratio, as calculated by the DMST code using experimental data as input [22].

S IMO F ERREIRA ET AL .
Ferreira developed a generalized objective function for VAWT airfoil optimization. The principles used come
from his previous work on VAWT wakes [46]. One important general remark in his paper is that with the
increase in computational power, more and more optimization algorithms are directly coupled with CFD
simulations. This approach lacks a clear rationale of what features an airfoil should have and is an inefficient
process. A good objective function which links the rotor performance to the airfoil performance could potentially speed up the process by orders of magnitude and increase the understanding of the aerodynamics
of VAWT [30].
The basic element of wake generation of the VAWT was taken as the starting point of the objective function.
Due to the azimuthal variation in bound circulation on the blade, vorticity is being shed constantly. Energy
extraction is directly dependent on the shedding of vorticity, so optimizing the shedding of vorticity will op@
@C c
timize energy extraction. This means optimizing @b which is proportional to @l . Translating this to airfoil
properties, the lift slope should be maximized to arrive at the maximum power.
This rationale was used by Ferreira to construct an objective function for an optimization. The optimization was performed using a genetic optimizer for airfoils developed by Oliveira [31]. In this optimization the
roughness sensitivity was incorporated by doing a clean and dirty simulation for each airfoil and averaging
the performance. Also a structural objective was added to increase the flapwise bending stiffness. This resulted in the AIR family of airfoils, which are relatively thick airfoils. Two examples can be found in figure
2.10.
These high thickness airfoils are beneficial for structural purposes. The reason these airfoils come out of the
optimization is that increasing thickness, up to a limit of about 35%, has no detrimental effect on the lift slope.
As a continuation of this research, one airfoil was tested in the low-turbulence wind tunnel of the TU Delft.
This airfoil was named DU12W262 [47] and can be found in figure 2.11.
Measurements were mostly taken with a PIV method which captures the flow field around the airfoil. They
were performed at Reynolds numbers around 1 million. The results concluded that the DU12W262 outperformed the NACA 4 series which were in the initial population of the genetic algorithm.

2.2.2. U NIVERSITY OF W INDSOR


At the university of Windsor in Canada, Islam conducted an investigation into desirable airfoil features for
VAWT [48] and deriving from this a special purpose VAWT airfoil [24]. The airfoil was designated for small
capacity VAWT which are characterized by airfoils operating at low Reynolds numbers in the order of 1 105 .
The main objective was to find the best performing airfoil out of a small group of six existing airfoils which are
suggested by literature. These airfoils are: NACA 0015, NACA 4415, GOE 420, NASA LS-0417, NASA NLF-0416
and the S1210.
The performance was calculated using a Cascade model, with correction modules for dynamic stall and flow

14

2. L ITERATURE R EVIEW

Figure 2.10: Two examples of the AIR family of airfoils [30].

Figure 2.11: A representation of the DU12W262 airfoil [47].

curvature. From this analysis, the NACA LS-0417 (or GA(W)-1, see figure 2.12) was found to be the best performing airfoil. This was the starting point of the airfoil design. Four aspects are taken into account in a
sensitivity analysis. The influence of camber, thickness, leading edge radius and trailing edge thickness are
taken into account and varied using XFOIL.
The resulting airfoil is the MI-VAWT1, see figure 2.13. The exact geometry is not known, but when comparing
it to its parent airfoil, the LS-0417, big differences are present. The MI-VAWT looks symmetric and also
shows a lot of resemblance to a NACA 0018 with a thinner tail.

Figure 2.12: A representation of the NASA LS-0417 airfoil [48].

2.2. R ECENT DEVELOPMENTS : 2005- PRESENT

15

Figure 2.13: A representation of the MI-VAWT1 airfoil [48].

2.2.3. N UMERICAL M ETHODS


The field of optimization has been influenced by the enormous increase in available computational power.
Heavy calculations involving panel methods or CFD calculations are becoming less expensive. This section
gives an overview of VAWT airfoil optimization campaigns conducted in this manner.
Carrigan designed a proof of concept optimization for VAWT airfoil during his master thesis in 2012 [25, 29].
Airfoil geometries were created by differential evolution from the NACA 4-series. This constrains the optimization to only this family of airfoils. A mesh generation module meshed the airfoil before analyzing it in a
CFD simulation in FLUENT. As a baseline geometry the NACA 0015 was used. Two optimization cases were
done. The first case was at a TSR of 1 and a rotor solidity of 1.5, which are both unusual values for VAWT. A
NACA profile with a maximum camber of 0.94% at 60% of the chord and a maximum thickness of 17.7% was
found (close to the NACA 1618 geometry). This airfoil showed a 2.4% increase in C p when compared to the
NACA 0015. The second case was run at a TSR of 3 and a solidity of 0.4, which is more toward normal VAWT
operating conditions. A 10.9% thick, symmetric airfoil resulted (close to the NACA 0011 geometry). This airfoil performs 1.1% better than the baseline of the NACA 0015. According to Carrigan, the symmetric airfoil
shape is due to the low solidity. With low solidity rotors, no blade-vortex interactions take place and positive
and negative angles of attack are of same magnitude.
In work of Vassberg et al. [21] CFD simulations are used to compute the performance of a modified version
of the NACA 0015, the WARP0015-RC8. This airfoil is designed using the same approach as Zervos [16], the
camber line warped along the circular path of the blade. This was done for a c/R of 0.125. Using the CFD
analysis, an increase in C p of 4% was observed.
Bourguet et al. [23] coupled the commercially available optimizer Optimus with the CFD tool StarCD. The
optimization was performed using three criteria: High nominal power production, high efficiency range and
blade weight. The optimization only took into account symmetric profiles. The resulting profile was very
close to the NACA 0025 and had a 28% increase in nominal power production and 46% increase in efficiency
range, compared to the NACA 0015 baseline design. It is not clear from the paper how this efficiency range
was measured.
Castelli [26] investigated the effect of relative thickness of the airfoil on the VAWT performance. Simulations
were done using the commercially available CFD software ANSYS FLUENT. NACA 0012 and NACA 0021 were
analyzed. It was found that the NACA 0021 had a 4.5% higher maximum C p at a lower TSR compared to the
NACA 0012. This was attributed to the fact that thicker airfoils stall at a higher angle of attack than its thinnner
counterpart.
Danao et al. [28] set up a CFD campaign to find the effect of airfoil thickness and camber on VAWT performance. Four airfoils were selected to do this investigation: the NACA 0012, NACA 0022, NACA 5522 and the
LS0421. It was found that the NACA 0012 performed better than the NACA 0022 for higher TSR.
Looking at all these campaigns it can be concluded that exploring full design spaces is still too computationally intensive. The demanding CFD simulations are mostly used to analyze effects of changing certain airfoil
characteristics or to check a theory. The most elaborate optimization of Carrigan was restricted to the NACA
4 series which limits the possible airfoil shapes considerably.

16

2. L ITERATURE R EVIEW

2.2.4. C OMMERCIAL DEVELOPMENTS


Several companies adopted the idea of offshore VAWT and are working on concepts and prototypes. A few
examples of current developments are given here. Not a lot of information is available, most observations are
made from visualizations of the concepts.
Gwind [49] is a high solidity helical H-type VAWT. The VAWT will be deployed on a floating structure
which will be stabilized by a gyroscope. At the end of 2013 the prototype Spinwind 1 was commissioned.
Estimating from the size this is a prototype in the order of 10 kW.
The Deepwind [50] project focuses on developing a -type floating turbine with a power rating of 5 MW.
Currently tests are being performed on a small prototype in the Roskilde Fjord in Denmark.
As part of the INFLOW project [51] Nenuphar [52] is developing a floating helical H-type VAWT. A 35
kW prototype was already installed and tested. The first stage of a 2 MW onshore prototype was built
in 2014, this first stage is rated at 600 kW.
Spinfloat [53] is developing 5 MW H-type VAWT also for a floating platform. The website mentions that
pitching blades will be part of the design, but so far no prototypes have been built.
VertAx Wind Ltd [54] is developing a bottom founded H-type VAWT with a rating of 10 MW. Wind tunnel
tests were performed on a VAWT but further information is not available.
All these companies are stating that the airfoils of the turbines are also subject of research, but since they are
ongoing commercial developments, nothing is shared. This overview serves as a reference to the landscape
of developments currently taking place.

2.3. C ONCLUSION
This review shows a summary of the research and design efforts into the airfoils for VAWT. In the early days the
main method to develop airfoils was to analytically derive principles which served as a basis for the design.
In this period mostly existing airfoils were used and then modified to get the desired performance. A clear
shift is shown in design methods with the arrival of more and more computational power. A trend is visible to
let the computer design the airfoil using combinations of optimization algorithms and aerodynamic analysis
codes, the theoretical/analytical basis is sometimes skipped. In my thesis I will try to speed up the process
by using an analytically derived function to do the optimization and then using more advanced software to
assess the performance of these airfoils and validating the analytical work.
Conclusions on possible objectives for VAWT airfoil design can be drawn on basis of this review. They will be
discussed in the list below.
@C

The lift slope @l is used by several references as an aerodynamic objective. In [12] it is used for achieving a high C l quickly, while Migliore et al. [14] and Ferreira et al. [30, 32] consider the lift slope in itself
the goal and high C l values are not required. The difference between these is that Migliore et al. discards the lift slope as an objective because too little variation can be observed between lift slope. This in
contrast to the findings of Ferreira et al. that found significant differences in power performance with
small changes in lift slope.
The drag is also an aerodynamic objective for many researches. It shows up in different forms: low C D,0
[15] and low C d in general [12, 14, 22, 30] are objectives that ensure the least negative torque from the
airfoils. However there might be a trade off when a small increase in drag can have beneficial results
for other objectives. This makes it a good secondary objective. Wide drag buckets are mentioned in
[15], which depends on the tip speed ratio of the turbine (TSR), at high TSR wide drag buckets are
not necessary. Finally, [12] opts that drag should be symmetric about the zero lift angle. This probably
comes from the assumption that the angle of attack distribution will be symmetric in up and downwind
regions of a VAWT, which is a faulty assumption.
Flow curvature is a phenomenon to be taken into account. Several researchers [14, 16, 21] have done
this by transforming the shape of an existing airfoil. By curving it along the streamlines the characteristics should become similar the the untransformed airfoil. By using CFD or some vortex models this
phenomenon is automatically taken into account, otherwise a correction or transformation should take
place to produce valid airfoil designs.

2.3. C ONCLUSION

17

The topic of roughness sensitivity is used as an objective by Sandia [17] and Ferreira et al. [32]. Soiling
of wind turbine blades can have a detrimental effect on power performance. The question that needs
to be answered is: What is more cost effective, designing blades that are roughness insensitive with
probable drawbacks in normal power output or cleaning wind turbine blades? This is dependent on
factors as the amount of reduction in performance with soiling, the degree of soiling and the cost of
cleaning. These factors are hard to predict, so a definitive answer to whether this design approach is
necessary cannot be given at this time.
A contrast can be seen between earlier research when most airfoils were designed to be symmetric, with
Healy [10, 11] concluding the closer the profile is to symmetric, the higher the output. Results of more
recent research show predominantly cambered profiles. Main reason for this is the recognition of the
different operation conditions between up and downwind halves of the rotor. Preset pitch and active
control are also ways to cope with this fact.
The airfoil design will be impacted by Reynolds number. The upscaling of wind turbines will also set
different requirements for airfoils. Most research has focused on small turbines with Reynolds numbers below a million. However, the developments mentioned in section 2.2.4 will be bigger turbines
operating at higher Reynolds numbers. This has to be factored in into the design process.
The key challenge for VAWT airfoil development at this time is to find a consistent (set of ) aerodynamic optimization objective(s) to be able to design airfoils for a rapidly developing and changing branch of industry.
This objective is ideally applicable to a range of VAWT scales, while factoring in phenomena like roughness
sensitivity and flow curvature.

3
A IRFOIL D ESIGN M ETHODOLOGY
In this chapter the theory for the optimization of VAWT airfoils will be explained. First an aerodynamic and
structural goal for an optimal airfoil will be derived in section 3.1. Section 3.2 will show how the phenomenon
of blade soiling will be incorporated into the design.

3.1. A IRFOIL OPTIMIZATION GOALS


The optimizer, which will be explained in section 4.1, works based on two objective functions. These calculate
the performance that the airfoils in the generated populations have. In this optimization an aerodynamic and
a structural objective function are used. These are usually conflicting objectives in wind energy, but also in
aeronautical engineering. For structural engineers it would be ideal to have thicker blades so the structural
elements could be designed more efficiently, however for aerodynamic engineers the thickness should be
limited in order to limit drag. Conflicting objectives are well suited for a Pareto optimization method. This
method will be explained in section 4.1.2. Both objective function will be explained in more detail in this
section.

3.1.1. A ERODYNAMIC F UNCTION


The energy extracted by a VAWT is a function of the location and strength of the wake over the boundary of
a control volume surrounding the VAWT [30]. The aerodynamic objective function works on the principle of
optimizing this wake shedding, the derivation of this work is cited from the work of Simo Ferreira [30].
According to Kelvins theorem, the VAWTs wake is generated by the temporal/azimuthal variation of bound circulation on the blade; its vorticity distribution is not dependent on the average
bound circulation on the blade but only on its azimuthal variation. Implicitly, the same power
conversion (same wake) can occur as the result of different loading distributions over the rotor/actuator surface. The optimization of the energy conversion can be performed by optimizing
the shedding of the wake. For a 2D VAWT defined by an actuator, the optimization of the wake
generation implies the optimization of the curl of the load distribution over the surface, leaving
undetermined a constant load distribution from the integral of the solution. However, an optimal loading distribution might in practice not be achievable or even desired. The question is as
follows: for a given wake distribution, which implies an azimuthal variation of bound circulation,
which aerofoil gives the most power converted into effective torque?
For a given number of blades and tip speed ratio , an azimuthal () distribution of shed vorticity
w () also implies a specific wake which induces a specific velocity field, resulting in a specific
azimuthal distribution of perceived velocity Vp () and inflow angle (). The distribution of shed
vorticity w () implies a distribution of the gradient of bound vorticity over the rotation
d ()

d (Vp C c)

d b ()
d .

We can define db = 12 d l since, for a given wake, the local perceived velocity distribution Vp () is defined, a distribution of the time-derivative (or alternatively azimuth-derivative) of

19

20

3. A IRFOIL D ESIGN M ETHODOLOGY


bound vorticity over the rotation is equivalent to implying a distribution of

dC l c
d .

dCc

Although there is a unique distribution d l that satisfies the generation of the wake, there is
an infinite number of suitable distributions C l c. For rotor design, these possible solutions can
be achieved by, for example, adding a fixed pitch angle , changing the aerofoil camber or other
dCc
forms of circulation control. Additionally, we can rewrite d l as a product of an average C l c and
a distribution function C () as
for a certain wake.

d Cl c
d

= C l c C () . The term C l c is not only a constant but specific

Ideally, we would convert the maximum of the energy extracted from the flow into mechanical
power associated to a torque; however, we are limited by the effect of drag. The average contribution of drag to torque (TD ) is defined by Equation 3.1, where R is the rotor radius, is the air
density, is the angle between the perceived velocity Vp and the tangent to the rotor at the blade,
and U1 is the unperturbed velocity upwind.
TD =

R2
0

R 12 cos Vp2C d c d = R 12 (U1 )2

with A 1 =

cos Vp2
(U1 )2

R2
0

A 1C d c d

(3.1)

For a given tip speed ratio, minimizing the impact of drag is equivalent to maximizing the inverse
of the integral in Equation 3.1. Fixing the variables that are not properties of the blade section,the
maximum performance can be expressed in Equation 3.2 as a maximum of the inverse of the
integral along the blade section design space variable .
d
d

R2
0

with

1
A 1C d c d

d2

R2
0

=0

1
A1C d c d

(3.2)
<0

Since C l c = const , Equation 3.2 can be rewritten as Equation 3.3


d
d

R2

with

C l c

=0

A1C d c d

C l c
d2
R
<0
2
d2
0 A1C d c d

(3.3)

To further simplify the analysis we define C d c as a product of the term C d c term and a distribution function G () , as C d c = C d cG () .
Equation 3.3 can then be simplified into Equation 3.4.
d
d

Cl

1
=0
C d 0 A 1 G () d

Cl
2
1
R2
with d 2
<0
d
C d 0 A 1 G () d
R2

(3.4)

Equation 3.4 gives the first insight into the aerodynamic optimization of a VAWT aerofoil: in a first
approximation, for a given range of angle of attack over the rotation, the aerodynamic optimizaCl

tion is achieved by maximizing . This result is analogous to the C l


optimum for horizontal
d max
Cd
R2
axis rotors. The term 0 A 1G () d is a weighing function which depends on the aerofoil polar
and should be accounted for.
For the case of no circulation control (excluding blade pitching), the azimuthal distribution C l
can be defined by

dCl d
d d ,

where is the angle of attack. As

and blade pitch angle distribution, the maximization of


Cl
Cd

Cl
Cd

d
d

is defined by the induction field

is equivalent to the maximization of

. We therefore arrive at an aerofoil aerodynamic objective function that is proportional to the

lift slope of the aerofoil C l and inversely proportional to drag.

3.2. R OUGHNESS MODELING

21

This derivation is important to the strategy of this optimization. The VAWT aerodynamics are used to derive
an objective on the airfoil scale, this means that for the optimization not every airfoil has to be simulated by
a complete VAWT model. Instead the airfoil can be assessed using simple tools (which will be explained in
section 4.1.3) which will save a lot of computational power. This means that the strategy can be implemented
such that big design spaces can be covered to find a good airfoil design.

Cl
Cd

(3.5)

max

As mentioned in the quote, the aerodynamic goal for VAWT as stated in expression 3.5 is analogous to the
HAWT optimization goal of glide ratio. The different goals originate in the fact that the VAWT is in unsteady
operation with constantly changing angles of attack while HAWT in principle work on fixed angles of attack.

3.1.2. S TRUCTURAL F UNCTION


For the structural objective there are several options, obvious choices would be edgewise or flapwise bending
stiffness. Edgewise bending is bending parallel to the chord and flapwise bending is bending perpendicular
to the chord. When thinking of wind turbines, fatigue can also be regarded as a big issue. However fatigue
depends very much on operating conditions and cannot be made into an usable objective for this optimization. The choice between the remaining two objectives is made in favor of flapwise bending stiffness. In this
direction the forces will be the biggest and the stiffness is generally smallest.
The structural objective is to maximize the geometrical stiffness of the airfoil. It is defined as the moment of
inertia about the horizontal neutral axis Itxxs , which corresponds to the flapwise bending stiffness of the VAWT
blade. The airfoil is assumed to be a thin-walled structure. This assumption makes the calculation simpler
and a skin thickness does not need to be specified. By maximizing the geometrical stiffness of the blade, less
material and stiffening structures will be needed to make the blade resistant to the loads. This will reduce cost
of the blade, but also reduce the weight, leading to less stringent requirements on other parts of the turbine
like the struts, tower, bearings etc.
In a real blade, stiffening structures will dominate the stiffness. Objective functions taking into account stiffening I-beams or box structures in the airfoil were investigated. This method was not adopted as the main
structural function because in this case the structure cannot be assumed thin-walled anymore, this means a
thickness or thickness ratio has to be chosen for the I-beams. This leads to an uncertain input because this
will depend very much on the shape of the final airfoil. It will however have a big impact on the outcome of
the structural function and thus have a big, uncertain influence on the optimization.

3.2. R OUGHNESS MODELING


As will be explained in more detail in section 4.1.3, airfoils operating in soiled conditions are taken into account in the optimization. This is done by modeling roughness on the airfoil. In RFOIL there are three options
which can be used to simulate roughness.
The Ncr i t amplification factor is a measure for the disturbance (turbulence) level of the flow. The e N
model for which it is used, can trigger the transition from laminar to turbulent flow. Ncr i t = 9 is used for
stable flow comparable to standard wind tunnel flow [55]. Rough conditions will trigger transition to
turbulent flow. The Ncr i t does not mimic roughness but mimics the effect of roughness. This is done by
placing the disturbances normally caused by roughness already in the flow, this makes the boundary
layer more susceptible to transition. This is a reasonable approximation to simulate roughness. By
setting the Ncr i t to a low value (below 1) roughness can be simulated.
The second tool is the manual forced transition point x t r , which can be set for both bottom and top of
the airfoil. By setting the transition point close to the nose of the airfoil, where usually the rough surface
is located, the flow will transition there always.
The airfoil soiling phenomenon needs to be studied in more detail to see which of these methods suits the
simulation. Figure 3.1 shows an example of blade soiling on a HAWT. The roughness can be caused by
squashed bugs, corrosion of leading edge and buildup of ice, salt or dust particles. The roughness is mostly
concentrated at the leading edge of the airfoil, the rough surface will make the flow transition from lami-

22

3. A IRFOIL D ESIGN M ETHODOLOGY

nar to turbulent there. The figure shows very severe soiling, normally the soiling is less and will thus impact
the transition less. There are two modeling approaches: 1) force transition at a certain point if roughness is
present, or 2) place disturbances in the flow to make it transition earlier.
If a low Ncr i t is used, the flow will transition close to the stagnation point, making the entire airfoil flow turbulent. A forced transition point models the physical phenomenon of roughness more accurate. But forcing
transition at a fixed point means that an estimate has to be made where flow should transition. A choice on
this position is made in section 4.1.3. Since Ncr i t mimics the effects of roughness on transition better than a
forced transition point, the Ncr i t factor is used a the primary simulation tool for roughness.

Figure 3.1: Example of blade soiling, bugs are squashed on impacting the blade [56].

To evaluate the impact of different soiling cases, simulations were done using the double wake model [67].
This is a VAWT simulation model which will be used to analyze the results of the optimization and will be
explained in section 5.2. Four cases were tested, one clean case as a reference and three soiled cases to see
the impact of surface roughness:
Severe soiling: N=0 (flow will transition almost instantly)
Medium soiling: N=4 and forced x t r = 0.1
Medium soiling: N=4 and forced x t r = 0.2
Clean case: N=9

Figure 3.2 shows the impact of these different soiling cases on the tangential force of a turbine, which is analogous to the performance. This comparison was made for a 26% thick airfoil. The clean case is shown by
the black line. It can be seen that the roughness cases all impact the performance, but the severity of the
roughness plays a role in the amount of performance degradation. In the upwind region (0 180 ) the influence is not that severe, in the downwind part (180 360 ) the influence is more pronounced. The roughness
causes separation in the upwind region and the vortices shed will be encountered in the downwind region.
Especially at 225 azimuth this impact is big.

3.3. O PTIMIZATION STRATEGY


The strategy to get to an optimized airfoil and a working optimization procedure is as follows. After the optimization tool (explained in chapter 4) has been tested and verified, a first set of inputs will be chosen to do a
first optimization. A choice will be made on the following properties/issues:
How to model soiling in RFOIL. This involves the choice of critical amplification factor Ncr i t and forced
transition points x t r .
Outer boundaries of optimization problem. These will be chosen to allow as much freedom as possible.
More stringent constraints will be defined in these boundaries.
Minimum trailing edge angle; this angle is set to avoid trailing edges which are to sharp for production.
Choice of parameters defining the operation conditions of the airfoil and VAWT. This involves choosing
a Reynolds number and an angle of attack range corresponding to the design tip speed ratio of the

3.3. O PTIMIZATION STRATEGY

23

Tangential Force for TSR = 4, pitch = 4


A26 N = 0
A26 N = 4, T=10
A26 N = 4, T=20
A26 N = 9

0.25
0.2

Tangential Force []

0.15
0.1
0.05
0
0.05
0.1
0.15
0.2
0

45

90

135

180
225
Azimuth []

270

315

360

Figure 3.2: Impact of roughness cases on tangential force for a 26% thick airfoil.

VAWT airfoil.
Find a suitable maximum thickness for the airfoils to avoid the inaccuracy region of RFOIL, but still
leave enough freedom for the optimization.
A suitable weighting between clean and soiled case must be determined. This will be investigated using
the first optimization trials.
From the results of this first optimization a few characteristic and interesting shapes will be chosen to do the
full VAWT analysis on. These shapes will be chosen not necessarily on basis of their best performance, but
also on their geometrical features. Seemingly weird geometries will be interesting to analyze to see if these
geometries are really favorable, or enter the optimization results on basis of other reasons. These reasons are
for instance inaccurate descriptions of some aerodynamic phenomena which result in undeserved higher
performances. Three airfoils are chosen for the full analysis and the results are used to choose the inputs for
a final optimization campaign. The results of these three airfoils are also published in [32].
The insight gained from this first optimization trial will be used to improve the optimization. The second
optimization trial will involve minor changes in the aerodynamic objective function and weighting between
clean and soiled performance. Four cases will be run, the results of these four optimizations will be analyzed
and the best way will be used as an example result of a good VAWT airfoil and this process will be recommended as the best way to optimize VAWT airfoils.

4
A IRFOIL G ENETIC O PTIMIZATION
Designing airfoils can be done in several ways, if there is already a general shape available for the application,
a popular way is to tune an existing airfoil to your needs. As can be concluded from the literature review in
chapter 2 no consensus has been reached on an airfoil shape for VAWT. This makes the possibilities of airfoil
tuning limited and other methods are needed. In this research a choice was made to investigate a big design
space in search for suitable airfoil shapes. This is done by using a genetic algorithm which generates airfoils
and evaluates them according to two objective functions. For this approach a lot of airfoils will need to be
aerodynamically simulated. To keep computational cost acceptable CFD methods are not feasible. To limit
computational cost the airfoil analysis program RFOIL is used. Because RFOIL simulates airfoils and not a
complete turbine, the challenge is to come up with a good objective function which will be able to predict the
performance of the wind turbine based on airfoil coefficients. To recall, the research goals of this thesis are:
To find an airfoil shape or family of shapes that is suitable for VAWT.
To validate the objective functions and optimization algorithm as an efficient optimization method to
design airfoils in future projects.
A program developed by Gael de Oliveira [31] called Optiflow will be used for the optimization. Optiflow is an
optimization algorithm for airfoils which is coupled to RFOIL for airfoil analysis. The work that needed to be
done consisted of choosing and verifying boundary conditions and inputs and implementing the objective
functions.
This chapter will elaborate on the optimization program in section 4.1. Next the constraints are explained in
section 4.2. The choice of constraints is important to the optimization because certain restrictions can have a
big impact on the final shapes. In section 4.3 the implementation is shown of the objective functions derived
in section 3.1.

4.1. O PTIMIZER
The optimizer Optiflow is built upon MATLABs multi criteria optimization tool gamultiobj which is coupled to work together with the airfoil analysis tool RFOIL. To facilitate the optimization the airfoils will be
parametrized. These facets of the optimization procedure will be explained here.

4.1.1. PARAMETERIZATION
Normally airfoils are represented by a set of coordinate points which describe the full shape of the airfoil in
(X,Y) coordinates. These are usually between 100 and 200 points. If this would be given to the optimizer as the
free parameters, the design space would be enormous. Next to this extra conditions should be implemented
to ensure smoothness of the airfoil. To be able to generate airfoil shapes efficiently, the airfoil shape is parameterized according to the Class Shape Transform (CST) method that was originally developed by Boeing [57].
The airfoil shape is represented by equation 4.1. In this equation there are three functions, the shape-function
S, the class-function C and the trailing edge function z t e , which are all functions of x.
25

26

4. A IRFOIL G ENETIC O PTIMIZATION

t ai r (x) = S(x)C (x) + z t e (x)

(4.1)

The class-function provides the basics of every airfoil shape. These basics are: 1) an infinite slope and finite
curvature at x = 0 to define the leading edge, 2) a finite value (zero) at x = 1 to define z t e and 3) a single
extreme in the [0,1] interval. The class-function is defined by equation 4.2 and a graphical representation is
shown in figure 4.1. This figure shows clearly why this function is a suitable class-function.
p
C (x) = (1 x) x

(4.2)

Class Function
1
0.8
0.6
0.4
0.2
0
0.2
0.4
0.6
0.8
1

0.2

0.4

0.6

0.8

Figure 4.1: Graphical representation of the class-function in blue and its negative in green.

The shape-function is the function that defines the airfoils specific shape. It is a Bzier curve, which consists
of a weighted sum of Bernstein polynomials, and is defined by equation 4.3. An example of the Bernstein
polynomials with weights (coefficients) of 1 is given in figure 4.2.

S(x) =

r =N
X1
r =0

s r,N (x)B n

!
N 1 r
s r,N (x) =
x (1 x)nr
r

(4.3)

(4.4)

The coefficients B n are used by the optimization as variables. Polynomials on both upper and lower side of
the airfoil are 8t h order, which leaves the optimizer with 16 coefficients to use in the genetic optimization
process. By varying these, all kinds of airfoil shapes can be generated and put to the test.
Finally the trailing edge function defines the thickness of the blunt trailing edge. Most airfoils have a finite
thickness at the trailing edge for production reasons and RFOIL convergence. The function used for this is a
simple linear function given in equation 4.5. This leads to a small addition of thickness over the whole airfoil,
but since the z t e is usually in the order of 0.1% chord it has no considerable influence. However, if in future
studies flatback airfoils with very blunt trailing edges are considered, this function needs to be adapted.
z t e (x) = z t e x

(4.5)

This last function introduces one more variable to the optimization next to the Bernstein coefficients. The z t e
can also be varied by the optimizer, normally it can be varied on the upper and lower side independently. To
keep the airfoils standard, both z t e are coupled so they will always be symmetric around the y=0 line. There
is a possibility to set z t e to a fixed value. This was tested but gave convergence issues in RFOIL as well as
problems in the optimizer itself. This constraint was therefor not used.

4.1. O PTIMIZER

27

Figure 4.2: Representation of the Bernstein basis polynomials that define the shape function [58].

The parameterization of the airfoil leaves the optimizer with a set of 17 variables (8 upper b n , 8 lower b n and
a z t e ). This amount is enough to do genetic optimization with. An example of airfoil coefficients is given in
table 4.1. These coefficients will produce the airfoil shown in figure 4.3.
Table 4.1: Example of coefficients for the airfoil AH93W257.

Upper side
Lower side
Trailing edge

0.3754
0.2377
0.0025

0.3731
0.3616
0

0.4185
0.2903

0.4432
0.3242

0.3063
0.3539

0.3535
-0.0828

0.1221
-0.0396

0.3028
-0.1392

0.3
0.2

y/c

0.1
0
0.1
0.2
0.3

0.2

0.4

0.6

0.8

x/c

Figure 4.3: AH93W257 airfoil generated from coefficients.

The upper side coefficients are generally higher than the lower side coefficients because the upper surface of
this airfoil is thicker. The last few coefficients on the lower side are negative to make the lower surface cross
the y = 0 line. The trailing edge is defined by only the first value, the second 0 is now a dummy variable which
is there just for if the trailing edge definition would change in the future.

4.1.2. M ATLAB S G ENETIC O PTIMIZER


MATLABs gamultiobj is based on the NSGA-II algorithm, which is a multi-objective genetic algorithm. The
algorithm works on basis of obtaining an optimal Pareto front. It also strives to maintain diversity in its

28

4. A IRFOIL G ENETIC O PTIMIZATION

population. This approach ensures that the optimization does not get stuck on local optima. To illustrate
the principle of the optimizer consider the following example. In figure 4.4, the blue squares indicate the
individuals of a population created in the optimization. The two axes give the values for the two objective
functions f 1 and f 2. In this example, both these values should be minimized to get the desired performance.

Figure 4.4: Example of Pareto front [59].

Consider three individuals in the population, individual A, B and C . When comparing A to B , A has a lower
(so better) score for f 2 and a higher (so worse) score for f 1. This means that both A and B have their positive
and negative aspects. However C has higher (worse) scores on both objective functions and is therefore dominated in performance by both A and B . The Pareto front will consist of individuals like A and B , the front will
be used to base the new population on.
Due to the nature of this optimization, the result will not be one single optimal airfoil. The result will be a
group of airfoils with a unique best combination of objective function scores. It will be up to the designer
to choose a suitable option for his application. In this optimization case the choice has to be made between
aerodynamics and structures. Generally a choice will be made more inclined towards aerodynamic performance. This is done because the applied aerodynamic objective functions inherently select thicker profiles.
Gains in aerodynamic properties are therefore considered more profitable.
A common way to validate optimization algorithms and to explain them in a simple way is to put them to the
test using the Schaffer tests. The first Schaffer test consists of the two objective functions shown in equation
4.6 and a plot can be found in 4.5a.

f or 3 x 5

f1 = x2

f 2 = (x 2)2

(4.6)

The Pareto optimum will be found for the interval x = [0, 2]. The real minimum is at x = 1, but since gamultiobj
does not decide which of the two objectives is more important, it finds an optimum range for the designer
to choose from. Points on the interval x = [0, 2] are comparable with points A and B of figure 4.4. If x = 3
is taken, this will correspond to point C since the values for both functions are higher than for instance for
x = 2.
The second schaffer test works with a parabola and a piecewise continuous function, which are shown in
figure 4.5b. The solution here is found in the intervals x = [1, 2] and x = [4, 5]. The second interval is there for
obvious reasons, the first interval x = [1, 2] is in the solution because the red (piecewise) function is lower than
anywhere else, which could be an optimum if the red function was deemed very important by the designer,
so it will show up in the Pareto. Gamultiobj was tested with both of these cases and proved to work.

4.1. O PTIMIZER

29

25

40
f1 = x2

f1 = piecewise function
35

f2 = (x2)
20

f2 = (x5)2

30
25

15
20
15
10
10
5

0
0
3

5
1

(a) Schaffer 1

(b) Schaffer 2

Figure 4.5: Plots of the two schaffer test objective functions.

4.1.3. RFOIL
For the aerodynamic assessment a program was needed that could deliver results fast with a reasonable accuracy. Full models of VAWT with CFD or vortex models were therefore out of the question, they would be to
computationally expensive. They will however be used to assess the final results of the optimization, more
on these methods can be found in chapter 5. To keep computational cost low, the choice was made for an
airfoil analysis tool instead of full VAWT simulation. This choice leads right to the XFOIL and RFOIL programs
which are known for their good combination of speed and accuracy.
RFOIL is an extended version of Mark Drelas XFOIL [60] and is used to compute airfoil coefficients. XFOIL
uses a 2D panel method with a strong viscous/inviscid interaction scheme [61] and is used extensively in the
aerospace industry to predict lift and drag coefficients. It is especially accurate at low angles of attack, at high
angles of attack separation will occur which cannot be modeled by RFOIL accurately.
With the development of RFOIL, the biggest improvement was made in the prediction of the lift around
C L,max [61], which makes the code much more suitable for the high angle of attack predictions common in
VAWT design. It was also designed to take into account rotational effects, which is useful for HAWT inboard
airfoils, but not relevant for the VAWT case.
To show the accuracy of RFOIL with respect to XFOIL and experimental data, a comparison is made in figure
4.6. A Reynolds number of 5 million is used and experimental data is taken from [62]. Since the objective
functions, which were discussed be explained in section 3.1, are mainly working in the linear part of the
C l curve, RFOIL gives a better prediction of the length of this straight part than XFOIL. The behavior in
stall is off, but this is less important for the optimization.
NACA 0018 for Re = 5 million
2

1.5

XFOIL
RFOIL
Experiments

Cl []

0.5

0.5

1
5

10
AoA [deg]

15

20

25

Figure 4.6: Comparison of C l curves of the NACA 0018 of XFOIL (blue), RFOIL (green) and experiments (red) for Re = 5 million.

30

4. A IRFOIL G ENETIC O PTIMIZATION

To be able to simulate roughness on the airfoil, RFOIL offers three tools, see section 3.2. The two tools that
will be used are restated here.
First there is the Ncr i t amplification factor, which is a measure for the stability of the flow and will trigger the transition from laminar to turbulent flow. By default RFOIL sets this factor to 9, which is a stable
flow comparable to standard wind tunnel flow [55]. In rough conditions, transition to turbulent flow
can occur very easily, so by setting the Ncr i t to a very low value (below 1) roughness can be simulated.
The second tool is the manual forced transition point x t r , which can be set for the both bottom and
top of the airfoil. By setting the transition point close to the nose of the airfoil, where usually the rough
surface is located, the flow will always transition there .
For the simulation of rough conditions on the airfoil, a combination of the two tools is opted for. It has to be
noted that with shifting the forced transition point forward, the Ncr i t becomes less important since the flow
transitions early anyway. The remaining question is where to force the transition of the flow. Several papers
suggest a very early transition on the upper side of the airfoil of about 1% and 10-15% on the lower side
[61, 63, 64]. However, since the airfoil is also simulated at high negative angles of attack, the flow will enter
the upper side of the airfoil after this 1% chordstation, so the flow does not transition. From a discussion with
Case van Dam from UC Davis [65] the transition points on both sides were chosen to be 10%. This is where
the soiling on the airfoil is normally located. For a more elaborate discussion of roughness modeling, the
reader is referred back to section 3.2.

4.2. C ONSTRAINTS
To make sure the optimizer comes up with feasible results, constraints are set within which the optimizer has
to operate. The constraints have different origins, some are there because of producability of the shape. Other
constraints have to do with the limits of analysis programs, in this case RFOIL. For certain thicknesses this
program will not produce trustworthy results. An overview of all the constraints will be given in this section.
The first constraints are set using a reference population of airfoils. This population contains existing wind
turbine airfoils and NACA 4 and 5 serie airfoils. From this population lower and upper bounds are set for the
airfoil pressure and suction sides. The actual bounds that are used in the optimization are shown graphically
in figure 4.7. The red part shows the part where the pressure side can be located, the blue side shows the
part of the suction side. In the purple area, both suction and pressure sides can be located. The reasons the
bounds look like they do are:
The bounds are set using a reference population and can then only be manipulated using factors to
grow or shrink the boundary.
The outer bounds are chosen such that the impact on airfoil shape generation is minimal.
At the front a white area is visible were no airfoil contour can be generated. This part is there because all
airfoils should have a thicker front part and this prevents the optimizer wasting time on configurations
in this region.
The purple region is used to give the airfoils a lot of freedom in the tail region. For instance the bottom
of the airfoil can take on positive y values in this region.
Boundaries have an impact on the optimization performance since they restrict the shape generation. The
impact of the bounds used in this optimization is that the leading edge radius is restricted to a range. This
range is set not to restrict leading edge radii that are expected. The maximum thickness dictated by the
bounds is higher than the maximum thickness constraint explained later in this section and thus has no influence. The impact of the bounds on the tail region is that the upper surface can not cross over into negative
y/c. This restricts certain shapes which could be interesting to VAWT. A recommendation is made in chapter
7 to alter the bounds generation to deal with this problem.
The purple area in figure 4.7 shows the need for another constraint, this constraint prevents the upper and
lower side of the airfoil from crossing eachother resulting in an impossible design. For this constraint the inequality constraint matrix is used provided by gamultiobj. This matrix prescribes boundaries for the Bernstein coefficients, in this case that B n,l o < B n,up (coefficients are both defined positive w.r.t. their airfoil
side).

4.3. I MPLEMENTATION OF OBJECTIVE FUNCTIONS

31

Figure 4.7: Example of global boundaries set for the airfoil optimization.

During the research the issue of very thin trailing edges came to light. These are difficult or impossible to
manufacture. Such shapes are excluded by imposing an inequality constraint on the trailing edge angle t ai l .
This is possible because of one of the fundamental properties of the parameterization, which is now derived.
To obtain the slope at the trailing edge of the airfoil, the derivative of the function describing the airfoil (equation 4.1) is taken and evaluated
at x = 1, see equation 4.7. From the properties described in section 4.1.1, it is

@C
known that C (1) = 0, @x
= 1 and S(1) = B N (see figure 4.2: only the last polynomial is 1 at x = 1).
x=1

@t
@S
@C
=
C
(1)
+
S(1)
+ zt e
@x x=1 @x x=1
@x x=1

(4.7)

To calculate t ai l the arctangent of the slope is taken, which will be approximated with a small angle approximation and result in equation 4.8. The minimum trailing edge angle is set at 8 deg.
up

lo
t ai l = B N + B N
2z t e

(4.8)

The final constraint that was added to the optimizer is a way to fix or limit the maximum thickness of the
airfoil. The limited thickness could not be implemented as a hard constraint yet, so a less elegant yet effective
method is used. In the structural objective function the thickness of the airfoil is checked. If the value exceeds
33% relative thickness, a penalty will be given to the airfoil, basically excluding it from the optimization. This
constraint was set because of the fidelity regions of RFOIL, which will start to give unreliable results at very
high airfoil thicknesses. Originally this was set at 30%, but to ensure freedom for the optimization algorithm
it was raised to 33%.

4.3. I MPLEMENTATION OF OBJECTIVE FUNCTIONS


The theory behind the objectives for the optimization is explained in chapter 3. The implementation of this
theory into workable objective functions is shown in this section.

4.3.1. A ERODYNAMIC OBJECTIVE FUNCTION


First an explanation will be given on how blade soiling is incorporated via a weighting function, then the step
by step approach to calculating the objective is shown.
Wind turbine blades are often subject to harsh environments. Dirt particles, like dust, insects or salt, can
accumulate on the blade making the surface rough and forcing transition of laminar to turbulent flow earlier
than in clean blade conditions. To account for this phenomenon, two airfoil simulations are made in RFOIL: a
simulation using a free transition of the flow and a simulation were the flow is forced to transition to turbulent

32

4. A IRFOIL G ENETIC O PTIMIZATION

around the nose of the airfoil. Both their performance is checked by the objective function and weighted
according to 4.9.

sc aer o =

1
wc

sc cl ean

(4.9)

+ sc w s

r oug h

The implementation of this theory into the optimizer objective function is done as follows:
1. The airfoil is simulated in RFOIL for both clean and soiled cases. For every optimization the soiling
conditions (Ncr i t and x t r ) can be specified. The relevant data (C l and C d for a range of ) from these
simulations is imported into the function.
2. A range of angle of attack is chosen which represents the TSR. An range of 15 degrees is used for a TSR
of 4-4.5.
C

3. The objective Cl is calculated for the clean (free transition) case. This is done using 2 filters, which use
d
the principle of convolution to find the part of maximum lift slope and the corresponding C d values.
The angle of attack at which this value is found will be recorded (this angle will be a measure for the
pitch angle).
4. Now the same is done for the soiled (forced transition) case, but at the optimum clean angle, so that
both cases will be evaluated at the same blade pitch.
5. To come up with a single objective function value the weighting shown in equation 4.9 is applied. The
weights in this expression can be varied with each optimization.
Cl
Cd

is calculated using filters, these filters use convolution to find the desired characteristics. The filter for the
lift polar finds the steepest part of the lift polar using a monotonous filter which takes into account the angle
of attack range specified. The filter for the drag polar uses a non-monotonous filter, the values of this filter
are based on the frequency of occurrence of an angle of attack. Figure 4.8 shows the angle of attack variation
during a rotation taken from an inviscid panel model. It can be observed that the frequency of occurrence
of the angle of attack around 6 is much higher than that of other angles. The C d at this lower end of the
spectrum is thus more important for the performance and has a higher weight.

12
10
8
6

AoA []

4
2
0
2
4
6
8

50

100

150
200
Azimuthal angle []

250

300

350

Figure 4.8: Typical angle of attack distribution over a rotation of a VAWT (TSR=4).

4.3.2. S TRUCTURAL OBJECTIVE FUNCTION


The airfoil is defined in a coordinate file that can be used by RFOIL. This coordinate file is also the input for
the structural objective function. From these coordinates the airfoil is divided into 5000 slices. With these
slices the neutral line is constructed. The final value of this function is then calculated by summing up all the
Steiner terms of the airfoil skin elements (assuming thin skin) around the neutral axis like in equation 4.10.
Definitions of this equation can be found in figure 4.9.

4.4. D ISCUSSION ON I NPUTS AND C ONSTRAINTS

I xx X
= dL y2
ts

33

(4.10)

Figure 4.9: Definition of symbols used in equation 4.10.

The structural function takes into account a correction for bad aerodynamic behavior. Since the optimization
is done using the Pareto front method, it is theoretically possible that a rectangular box could show up as the
optimal structural solution. Since the optimization aims for airfoils, the aerodynamic objective function value
is used to filter out these unpractical solutions. If the aerodynamic performance is below a certain threshold,
the design gets a penalty and is dropped from the Pareto. This threshold is set at about half the maximum
aerodynamic performance. As discussed earlier, the same thing happens with airfoils above 33% relative
thickness (t/c).

4.4. D ISCUSSION ON I NPUTS AND C ONSTRAINTS


To carry out the optimization with success, the optimizer needs to be provided with inputs and certain
boundary conditions (constraints). This section will summarize the inputs and constraints and discuss the
effect they have on the outcome of the optimization.
Reynolds number: The Reynolds number of 5 106 is representative for large scale VAWT. A reference
simulation performed on a 5 MW vertical axis wind turbine showed a variation of Reynolds number
between 3 106 and 8 106 at a TSR of 4 and a wind speed of 10 m/s. At these high Reynolds numbers
the behavior of the airfoil is not changing that much with Reynolds number anymore, which makes
Re = 5 106 a good approximation.

RFOIL soiling simulation: The clean case is taken as free transition and a Ncr i t of 9. This is common
and recommended in the XFOIL manual [55]. For the soiled case no real research has been done on how
to model it. To come to a conservative conclusion a worst case approach was used which forced transition over the full airfoil. This conservative method will output airfoils with good soiling performance
in any case.
Trailing edge angle: The minimum trailing edge angle (t ai l ) is set at 8 for production reasons. Thin
trailing edges have an advantage of lower drag but are not feasible to build.
Maximum thickness: The maximum relative thickness of the airfoil was 33% (or 0.33 t/c). This has
both to do with the limits of RFOIL and the bad performance of thick profiles in soiled conditions. This
maximum thickness was extended to provide the optimizer with enough freedom.
General boundaries Figure 4.7 shows the general boundaries of the airfoil in the optimization. These
were taken as wide as possible to not have an influence on the outcome of the optimizer. Especially the
tail shape is important to give freedom.
Angle of attack range: The angle of attack range of 15 that is used to calculated the aerodynamic
performance of the airfoil is representative of the TSR. Figure 4.8 shows a typical variation of angle of
attack for TSR=4. This bandwidth is 18 but since the aim is TSR 4-4.5 the range used is 15 .

5
A IRFOIL P ERFORMANCE A NALYSIS
In order to verify the objective function and check performance, the results of the optimization are analyzed
using different methods for full VAWT simulation. All the results of the optimization are simulated using
an inviscid panel method, which is explained in section 5.1. Cooperation with third parties also provided
verification information, these methods are briefly explained in sections 5.2 and 5.3. A comparison between
the 3 analysis models will be given in section 5.4.

5.1. I NVISCID PANEL M ODEL


For quick analysis of the optimization results an inviscid 2D panel model [46] is used to analyze performance
of the airfoils in a VAWT situation. This method is used because fairly reliable results can be obtained at
low computational cost. This quick way will give insight right after the optimization in how the airfoils will
perform. In the panel model a simplified wind turbine is modeled in 2D. Only the airfoils are modeled, the
tower and struts, but also tip effects are not modeled. This turbine is simulated at a certain wind speed at
different TSR ranging from 2.5 to 7. The solidity is defined as B2Rc and ranged from 0.07 to 0.14. These ranges
are used to get a good overview of performance for many operating conditions. The TSR range covers all
the whole range for normal operation, while the solidity covers a range around the design solidity of 0.1. A
turbine with 2 blades is used. The airfoil that is used is a NACA 0015 for all simulations, the explanation for
this will follow later in this section.
In the panel model the airfoil is discretized with source and doublet panels. A Dirichlet boundary condition will ensure no flow through the panels (airfoil wall). The airfoil experiences a combination of velocities
from the wind, rotation, other airfoils and an induced velocity by the wakes vortices. This will determine the
strength of the doublet/source panel. At the trailing edge the Kutta condition is imposed, fixing the stagnation point on the trailing edge. The near wake is modeled by doublet panels, while the rest of the wake is
modeled by vortex points. Figure 5.1 shows a representation of the model.

Figure 5.1: Representation of panel distribution over airfoil and in wake (figure adapted from [66].

An example of the structure of the vortices in the wake is given in figure 5.2. On the left side the airfoils rotate
around point (0,0). Wind is moving from left to right, convecting the vortex points downstream. Interesting
features are seen at both edges of the wake where most blade-wake interaction takes place and at the core of
the wake where the induction velocity is most visible (vortex points contract).
This panel model is inviscid and will not give correct results on blade performance since no drag is present in
this model. It does however give fairly good results of the wake structure and induction on the blades. For this
35

36

5. A IRFOIL P ERFORMANCE A NALYSIS

40
30
20
10
0
10
20
30
40
50

50

100

150

200

250

300

Figure 5.2: Structure of the wake as simulated by the panel model after 5 rotations for TSR=3.

reason the angle of attack distribution and relative velocity distribution are extracted from this model. These
distributions are based on the wake structure and, together with airfoil data from XFOIL/RFOIL, are used to
calculate performance of airfoils. An example of the results used for this calculation is shown in figure 5.3.

10

4.8

4.6

4.4

4.2
U/U

AoA []

12

2
0

4
3.8

2
3.6
4
3.4
6
8

3.2
0

50

100

150
200
Azimuthal angle []

250

300

350

45

(a) Angle of attack distribution for TSR=4.

90

135

180
225
Azimuth []

270

315

360

(b) Relative velocity distribution on airfoil for TSR=4.

Figure 5.3: Results used from panel model.

The results of the panel model are produced for the NACA 0015 airfoil for different TSR and solidities of the
turbine. Other airfoils will behave differently than the NACA 0015, to correct for this, the solidity is corrected
using the lift slope. This is done by taking the ratio of lift slope of the NACA 0015 and the optimized airfoil
and multiplying it with the solidity, like in equation 5.1. The results for the panel model were calculated for
the range of solidity from 0.07 to 0.14. The corrected solidity is matched to one of these results. This is done
to make the inviscid loading equivalent to the viscous loading of the optimized airfoil, so the correct velocity
and and angle of attack distribution will be used. Since in inviscid flow, no separation is present, the (viscous)
optimized airfoil will be penalized with a higher loading beyond the stall angle. This penalty consists of a
higher induction when the airfoil reaches high angles of attack and relates to the model inviscid formulation.

cor =

@C l
@ 0015

@C l
@ OP T

(5.1)

Another correction is added which incorporates flow curvature in the model. A simple assumption is made
that this phenomenon will add to the angle of attack. This angle will depend solely on the solidity of the
turbine (c/R). To give an idea of the impact of this correction, a solidity of 0.1 will give a flow curvature angle
of 2.8 .
The optimal pitch angle is also determined by trail and error, pitch angles between -5 and 2 are tested. This
range is where the optimal pitch is likely to be. This is derived from the angle of attack distribution for one
rotation. The pitch that results in the highest C P will be selected.

5.2. D OUBLE WAKE PANEL M ODEL

37

5.2. D OUBLE WAKE PANEL M ODEL


To further investigate the airfoils behavior, other methods are also used to assess performance. The method
discussed in this section is an extension of the panel model which incorporates stall. This double wake panel
model, developed by Zanon [67], solves the unsteady potential flow panel equations together with the integral
boundary layer equations. At the separation point, an extra panel is introduced to model the separation of
flow. A schematic of this model is shown in 5.4.

Figure 5.4: Representation of panel distribution over airfoil and in wake [66].

The model is based on that at separation, vorticity is continuously shed along the shear layer between the
flow and the separation bubble. The location of separation is predicted by the boundary layer model. The
boundary layer model uses Drelas integral approach for steady flows [60], which is based on von Karmans
momentum equation and shape parameter equation. For the laminar-turbulent transition, the model uses
the simplified version of the e N method.
The separation criterion used in the present model is based on the value of the skin friction coefficient, C f
(flow separates when C f becomes negative). This criterion has been shown to be reliable in previous works
[68], but it fails when applied to pitching airfoils at medium-low Reynolds numbers, because it does not allow
flow reattachment after deep stall at the leading edge has occurred. An improved criterion is implemented in
the current model to identify the condition under which the turbulent reattachment becomes possible. This
criterion and the treatment of the reattachment process are described in detail in [66].
The double-wake model is able to capture the complex evolution of the vorticity field associated to the energy
conversion process of the VAWT. Since the model takes into account a viscous boundary layer, the predicted
values of the forces can be used directly from the model without interference of airfoil coefficient generation.
The force results are comparable to the ones provided by the much more expensive CFD approaches. The
double-wake model was validated in [66] with reference to steady and pitching airfoils NACA 0015 and NACA
0012 and to a VAWT in dynamic stall.
This model requires airfoils with closed trailing edges in order to work. The airfoils that result from the optimization will have a small trailing edge gap. This is standard in XFOIL/RFOIL and helps the convergence,
also for production reasons, the trailing edge cannot be knife-sharp. To make the airfoils compatible with
this model, the trailing edge is closed. This is done only for this model, so the final shape will still have the
open trailing edge. This will have the implication that a different shape will be analyzed by the double wake
model than the result from the optimization. To see the influence of closing the trailing edge both airfoils are
simulated in RFOIL, the result is given in 5.5.

5.3. CFD S IMULATION


NALU is a low-Mach number computational fluid dynamics code that can be configured to solve the incompressible Navier-Stokes equations using various turbulence model options including RANS, LES, and hybrid
RANS-LES models. The code uses an unstructured mesh algorithm with a fully implicit solver designed and
demonstrated to efficiently scale on many-core high-performance computing clusters. NALU allows for nonconforming mesh interfaces, which has been exercised in the present simulations to create a circular rotating
mesh block containing the VAWT rotor. In the current simulations, the SST-DES hybrid RANS/LES turbulence
model has been applied, without any explicit model for boundary layer transition. In this way, at present, the
CFD model provides a prediction for worst case boundary layer transition, where the boundary layer is im-

38

5. A IRFOIL P ERFORMANCE A NALYSIS

2.5
Open TE
Closed TE
2

1.5

0.5

0.5

1
10

10

15

20

Figure 5.5: Difference in lift curve for open and closed trailing edge, similar differences are found for drag.

mediately turbulent very close to the leading edge. An example of a resulting vorticity field from the CFD
code is found in figure 5.6.

Figure 5.6: Example of a vorticity field calculated by the NALU CFD model ( = 4).

5.4. C OMPARISON BETWEEN MODELS


In this section a comparison is made between the three models used in the analysis of the airfoils on a VAWT.
For this comparison one of the first results of the optimization is used, the RK1-26. This airfoil will be discussed in more detail in chapter 6. The results shown here are only used to compare the models.
First the performance of the models in the clean case will be shown. In this comparison the results from
the CFD method are missing since, due to the nature of its turbulence model, it can only simulate soiled
conditions. Figure 5.7 shows the normal and tangential forces on a VAWT blade for one rotation at a tip speed
ratio of 3. It is clear that there are quite some discrepancies between the two models, especially in the upwind
region. The cause of this difference is that stall is not well modeled in the inviscid panel model. The upwind
part of the rotation experiences high angles of attack and dynamic stall will occur at this TSR.
If the TSR is increased to 4, the differences become less as seen in figure 5.8. Due to the lower angles of attack,
there is less stall and the two models are agreeing more. The difference in height and position of maximum
has probably to do with dynamic stall, which is captured in the double wake model. Dynamic stall raises the
lift coefficient temporarily before dropping it quicker, which describes what is seen in the figure.
When looking at the soiled cases, the expectation is to see discrepancies between the double wake and the
inviscid panel model. This is because in soiled conditions, the airfoil will stall earlier which is not dealt with

5.4. C OMPARISON BETWEEN MODELS

39

Normal Force for TSR = 3 and N = 9


2.5

Tangential Force for TSR = 3 and N = 9

Inviscid panel model


Double wake panel model

0.5
Tangential Force []

1.5
Normal Force []

Inviscid panel model


Double wake panel model

0.6

0.5

0.4

0.3

0.2

0
0.1
0.5
0
0

45

90

135

180
225
Azimuth []

270

315

360

45

(a) Normal force

90

135

180
225
Azimuth []

270

315

360

(b) Tangential force

Figure 5.7: Forces on the RK1-26 for TSR=3 for 1 rotation in clean conditions.
Normal Force for TSR = 4 and N = 9
2
1.8
1.6

0.3

1.4

0.25

1.2
1
0.8

0.2
0.15

0.6

0.1

0.4

0.05

0.2

Inviscid panel model


Double wake panel model

0.35

Tangential Force []

Normal Force []

Tangential Force for TSR = 4 and N = 9

Inviscid panel model


Double wake panel model

0
0

45

90

135

180
225
Azimuth []

(a) Normal force

270

315

360

45

90

135

180
225
Azimuth []

270

315

360

(b) Tangential force

Figure 5.8: Forces on the RK1-26 for TSR=4 for 1 rotation in clean conditions.

correctly by the inviscid model. Figure 5.9 and 5.10 show that this expectation is right. The inviscid panel
model is way off the other two models predictions. It is interesting to see that the double wake model and
the CFD model agree well. Only figure 5.8b shows a discrepancy between double wake and CFD models, this
probably has to do with the generation of results of the two models. The double wake model averages the
results over 5 rotations while the CFD uses just one rotation. This means CFD can have more oscillations
and sudden influences. In this case the blade probably crosses a vortex core which increases lift suddenly,
creating extra torque.
The CFD model shows more oscillations than the double wake model. This is not the case, the double wake
results are averaged over several rotations, filtering out the high frequency oscillations. The results of both
models seem to follow the same trend in behavior. They are less predictable in the downwind part of the
rotation in which most of the blade wake interaction takes place. However the two models do catch the same
trends in their simulation.
Due to this good agreement between CFD and double wake model, the latter one is chosen to do the principle
full VAWT analysis with. The double wake model is computationally cheaper than the CFD and gives similar
results. The first analysis will still be done using the inviscid panel model because of its speed (in the order of
minutes, compared to the hours of the double wake model). This model will give a good first impression on
which airfoils to choose from the optimization, especially for higher TSR where less stall is occurring.

40

5. A IRFOIL P ERFORMANCE A NALYSIS

Normal Force for TSR = 3 and N = 0

Tangential Force for TSR = 3 and N = 0

Inviscid panel model


Double wake panel model
CFD simulation

2.5
2

0.3
0.25
Tangential Force []

1.5
Normal Force []

Inviscid panel model


Double wake panel model
CFD simulation

0.35

1
0.5
0

0.2
0.15
0.1
0.05
0

0.5

0.05
1
0.1
0

45

90

135

180
225
Azimuth []

270

315

360

45

(a) Normal force

90

135

180
225
Azimuth []

270

315

360

(b) Tangential force

Figure 5.9: Forces on the RK1-26 for TSR=3 for 1 rotation in soiled conditions.

Normal Force for TSR = 4 and N = 0


2

Tangential Force for TSR = 4 and N = 0


0.25

Inviscid panel model


Double wake panel model
CFD simulation

Tangential Force []

Normal Force []

1.5

0.5

Inviscid panel model


Double wake panel model
CFD simulation

0.2

0.15

0.1

0.05

0
0.5
0.05
1

0.1
0

45

90

135

180
225
Azimuth []

(a) Normal force

270

315

360

45

90

135

180
225
Azimuth []

(b) Tangential force

Figure 5.10: Forces on the RK1-26 for TSR=4 for 1 rotation in soiled conditions.

270

315

360

6
R ESULTS OF O PTIMIZATION
In this section the resulting airfoils of the optimization and the performance of these airfoils computed using
the VAWT analysis codes will be presented. The first optimization strategy was set up to find the influences of
using specific goals, constraints and weightings in section 6.1. Four other strategies were implemented, one
of which resulted in a final airfoil. These strategies and final airfoil will be presented in section 6.2.

6.1. O PTIMIZATION S TRATEGY 1


The first optimization strategy aimed for an airfoil robust to soiling, results were generated and three airfoils
were selected. With these results the full optimization process and analysis procedure can be run to see if it
all works and to see if the objective functions and boundaries are valid. Analysis will be done only with the
inviscid panel model and the double wake panel model. CFD results will not be used because section 5.4
showed that these results are similar to the double wake panel model, running both analysis wouldnt justify
the computational cost involved. This section will show the results of the first strategy and a discussion of
these results. This discussion is used to facilitate changes and improvements to the optimization process.
Three airfoils were taken from the optimization, these are called the RK1 series airfoils (1 for first generation).
Airfoils of thicknesses 23, 26 and 32% are analyzed.

6.1.1. I NPUTS OPTIMIZATION


The inputs for the first optimization strategy consist of inputs for the optimization program, RFOIL and the
objective functions. For the RK1-26 and the RK1-32 the same inputs were used. For the RK1-23, which was
a very early optimization, different inputs are used. All three airfoils were optimized for an angle of attack
range of 15 , which is representative of a TSR of 4.
The RK1-23 was optimized for a Reynolds number of 1 106 . For the aerodynamic coefficients, the clean case
was taken as N=9 with free transition and the soiled case was taken as N=0 with forced transition x t r at 0.05
x/c on both sides. The objective functions weights (equation 4.9) for clean and soiled (w c and w s ) were both
taken to be 0.5. In the aerodynamic function penalties were given for exceeding the maximum thickness of
0.33 t/c and trailing edge angle of 8 . Since this was one of the first runs, the trailing edge angle constraint did
not function properly yet.
The RK1-26 and RK1-32 were optimized for a Reynolds number of 5 106 , which corresponds to large VAWT
(> 5MW ). After research and discussions with experts1 , roughness was simulated using N=0 and x t r = 0.1 on
both sides. This represents a worst-case scenario for soiling and is believed to result in the most robust airfoil
shape against soiling. Weights between soiled and clean conditions were kept at 0.5. This was done to see
what influence this weighting would have on the soiled performance of the airfoil. This was updated after the
first results were analyzed. The trailing edge angle was taken out of the objective function and set as a hard
constraint and the maximum thickness was taken out to ensure freedom in the optimization.
1 Conversations with Gael de Oliveira and Alessandro Zanon

41

42

6. R ESULTS OF O PTIMIZATION

6.1.2. RK1-23
The RK1-23 is a 23% thick airfoil with a slight positive camber. The trailing edge is very thin, due to a disfunctional trailing edge penalty. The penalty was wrongly inserted and was basically turned off. The airfoil
is shown in figure 6.1, the coordinates can be found in appendix A. The RK1-23 was the best aerodynamic
performer according to the pareto. The tail shape of the airfoil results in extra loading on the tail of the airfoil.
This increases lift coefficient but also the (negative) pitching moment of the airfoil (for positive ). In the
way the airfoil will be mounted, this moment will counteract the torque of the turbine. The extra pitching
moment will also set extra requirements for the structural design of the blade.

0.2
0.15
0.1
x/c 0.05
0
0.05
0.1
0.15
0.2
0.1

t/cmax=0.23

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

1.1

y/c
Figure 6.1: Geometry of RK1-23 airfoil.

The performance of the RK1-23 is shown in figures 6.2a and 6.2b. This performance was calculated using
the inviscid panel model combined with the viscous RFOIL polars. This means that the effects of stall and
separation are only partly evaluated. The C P of the airfoil is highest around = 3.5 to = 4, which is the range
it was designed for. The reduced performance in soiled conditions can mainly be contributed to increased
drag and earlier stall, as can be seen in figure 6.2c and 6.2d.
A note on the shape of the C P , plots. With increasing solidity the maximum power shifts to a lower TSR. This
is the case for all airfoils simulated with the inviscid panel model. The reason for this is that increased solidity
will generally result in higher torque, this means the same power can be achieved at lower rotational speed.
This leads to the characteristic shape of these contour plots.
The analysis of the RK1-23 was continued with the double wake panel model. This model should incorporate
the separation of flow much better and give more insight into the dynamic behavior of the airfoil. Due to substantial computational time, only a solidity of 0.1 is simulated using the double wake code. The simulations
were run for a TSR of 2, 3 and 4. The C P is shown in figure 6.3, there is difference in the prediction of both
codes. The main conclusions are that soiling of the airfoil reduces performance and that the inviscid panel
model is conservative at higher TSR w.r.t. the double wake model.

6.1.3. RK1-26
The RK1-26 is a 26% thick airfoil with positive camber. The airfoil is shown in figure 6.4, the coordinates can
be found in appendix A. This airfoil is representative for a lot of solutions from the optimization. This family
has the main features of a thick front part of the airfoil where most of the structural stiffness is achieved,
combined with a highly loaded tail. This loaded tail increases the lift slope, because the lift increases faster
on the tail. This shape also comes with a nose-down pitching moment.
The performance of the RK1-26 is shown in figure 6.5a and 6.5b. Compared to the RK1-23 the clean performance is similar, when looking at the soiled case, the performance is somewhat lower than the RK1-23.
However it is again clearly seen that soiling on the airfoil will impact the performance a lot.
The double wake model gives more insight into the behavior of the RK1-26. In figure 6.6 it is also clearly visible
that the performance of the RK1-26 in soiled conditions drops drastically. Because of the cambered profile of
the RK1-26, it was found that using the airfoil at a fixed pitch angle would increase the performance. For the
RK1-26, this pitch angle was found to be around = 4 . This performance increase is mainly achieved due
to the avoidance of stall. For the rest of the analysis the optimum case with pitch is used.

6.1. O PTIMIZATION S TRATEGY 1

0.13

0.35
0.49

0.12
0.11

43

0.47 0.45

0.4

0.13

0.1
0.15
0.2

0.25

0.3

0.3

0.12

0.51
0.4

0.35

0.09

0.49

0.06

0.4
0.4

0.25
0.3
0.4

0.06

0.35

0.45

0.4

0.45

0.05

0.05
0.25

0.4

4.5

5.5

6.5

0.04
3

0.2

0.3

3.5

0.35

4.5

5.5

6.5

10

15

20

(b) C P (,) soiled case.


2
1.75
1.5
1.25
1
0.75
0.5
Cl 0.25
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
20

N=9 free transition


N=0 forced transition at x/c = 0.1

0.04

(a) C P (,) clean case.

0.02

0.35

0.07 0.3 0.35

0.47
0.4

0.1
0.2 0.15

0.3

0.08

0.35
0.4

0.47
0.45

2
1.75
1.5
1.25
1
0.75
0.5
Cl 0.25
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
0

0.25

0.09

0.47 0.45

0.49

0.35
0.04 0.3
3
3.5

0.15

0.1

0.25
0.3

0.07

0.2

0.11

0.1

0.08

0.10.05

0.25

0.35

0.4

0.06

0.08

0.1

Cd

N=9 free transition


N=0 forced transition at x/c = 0.1

15

10

( )

(c) C l C d

(d) C l

Figure 6.2: Performance simulation results for the RK1-23 airfoil from the inviscid panel model.

RK123 CP
0.6

0.5

0.4

0.3

0.2

Inviscid clean
Double wake clean
Inviscid soiled
Double wake soiled

0.1

2.5

3.5

4
TSR

4.5

5.5

Figure 6.3: C P T SR of RK1-23 airfoil for a solidity of 0.1.

6.1.4. RK1-32
The RK1-32 airfoil is shown in figure 6.7, the coordinates can be found in appendix A. This airfoil is a very
thick airfoil with positive camber. The main feature of this airfoil is its bulging tail shape. This shape was
found on the part of the pareto front which is more inclined towards optimal structural performance. The
tail shape reminds of an airfoil with a blunt trailing edge, with the current optimization a sharp trailing edge

44

6. R ESULTS OF O PTIMIZATION

0.2
0.15
0.1
y/c 0.05
0
0.05
0.1
0.15
0.2
0.1

t/cmax=0.26

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

1.1

x/c
Figure 6.4: Geometry of RK1-26 airfoil.

0.13
0.12

0.2

0.35
0.51

0.49 0.47

0.45

0.15

0.3

0.11
0.35

0.51

0.49

0.1

0.25
0.3

0.4

0.49

0.05
0.04
3

0.3
0.25

0.08

0.47
0.45

0.35

0.2

0.07
0.4

0.4

0.47
0.45

0.35
0.3

4.5

0.25
0.35

0.06
0.2

0.45

0.05

0.25

0.3

4.5

5.5

6.5

0.04
3

3.5

0.06

5.5

6.5

10

15

20

(b) C P (,) soiled case.

N=9 free transition


N=0 forced transition at x/c = 0.1

0.04

0.3

(a) C P (,) clean case.

0.02

0.35

0.15

0.4

3.5

0
0.05
0.1
0.15

0.3

2
1.75
1.5
1.25
1
0.75
0.5
Cl 0.25
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
0

0.25

0.35

0.35

0.09

0.47 0.45

0.07
0.06

0.2

0.2

0.1

0.08

0.1
0.15

0.3

0.12 0.3

0.11

0.09

0.25

0.35

0.13

0.1

0.25

0.4

0.08

0.1

Cd

(c) C l C d

2
1.75
1.5
1.25
1
0.75
0.5
Cl 0.25
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
20

N=9 free transition


N=0 forced transition at x/c = 0.1

15

10

( )
(d) C l

Figure 6.5: Performance simulation results for the RK1-26 airfoil from the inviscid panel model.

was required but in future research the possibility of blunt trailing edges could be investigated. Still as figure
6.8a shows, the aerodynamic performance is fairly good when analysed with the inviscid panel model. Only
in soiled performance (figure 6.8b) the performance deteriorates significantly. Especially for the design-TSR
of 4 the difference in performance is big, dropping from a C P of 0.45 to 0.2.
The double wake panel model agrees with this conclusion, the clean performance is remarkably good for
such a thick airfoil, but soiled performance is low (inviscid) or even below zero (double wake).

6.1.5. D ISCUSSION OF STRATEGY 1


This first family of airfoils was analyzed using the three analysis codes described in chapter 5. Both panel
models agree on the high aerodynamic performance of all three airfoils in the clean case. Figure 6.10 shows
the coefficient of performance gotten from both models compared with the NACA 0018 performance. It is

6.1. O PTIMIZATION S TRATEGY 1

45
RK126 C

RK126 C

0.6

0.6
0.5
0.5
0.4

CP

CP

0.4
0.3

0.3
0.2
0.2

2.5

3.5

4
TSR

4.5

5.5

Inviscid clean
Double wake clean
Inviscid soiled
Double wake soiled

0.1

Inviscid clean
Double wake clean
Inviscid soiled
Double wake soiled

0.1

2.5

3.5

(a) C P T SR of RK1-26 airfoil for a solidity of 0.1, no pitch.

4
TSR

4.5

5.5

(b) C P T SR of RK1-26 airfoil for a solidity of 0.1, with 4


pitch.

Figure 6.6: C P T SR of RK1-26 airfoil for a solidity of 0.1.

0.2
0.15
0.1
x/c 0.05
0
0.05
0.1
0.15
0.2
0.1

t/cmax=0.32

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

1.1

y/c
Figure 6.7: Geometry of RK1-32 airfoil.

interesting to see that much thicker airfoils can attain even higher efficiencies than the NACA 0018 airfoil. It
can be concluded that the objectives for the optimization result in improved airfoil performance compared
to NACA 0018 in clean conditions. Despite the 50/50 weighting of clean and soiled cases, the airfoils are sensitive to roughness with aerodynamic performance dropping significantly. Keeping in mind that the most
severe roughness case (N=0) was used. So all conclusions made in this section will change if the use of a less
stringent case of roughness modeling is deemed more suitable for actual soiling modeling. In next optimizations the weighting of the clean and soiled conditions should be more in favor of the soiled condition to get
better performance in these conditions. A fully soiled optimization will also be done, this will make a clean
simulation in RFOIL no longer necessary which will reduce the computational time for the optimization to
about half. The results of this will be shown in section 6.2.
A second conclusion that can be drawn from these first results is that the inviscid panel model (combined
with RFOIL data) will overestimate the soiled performance. This happens because separation is more common in soiled conditions, and RFOIL models the influence of this on the lift and drag coefficients only partly.
For higher TSR and clean conditions the difference between the two models is smaller and the inviscid panel
model can be used with more confidence to predict performance. This is because less separation and vortex
shedding is present in these conditions. Figures showing this can be found in figures 5.7 and 5.8 in section
5.4.
Before the optimization began it was assumed that the movement of the transition point in clean conditions
should give an indication of the degree of roughness sensitivity of the airfoil. Results from the inviscid panel
model show that the NACA 0018 is least roughness sensitive followed by the RK1-23, then the RK1-26 and
finally the RK1-32. The movement of the transition point in clean conditions can be seen in figure 6.11. At
first sight it looks like the assumption was incorrect, the overall transition point has an opposite trend to
the results of the inviscid panel model. But transition has a negative effect mainly when it induces earlier
separation. This earlier separation only happens on the suction side of the airfoil, so the view has to be
restricted to an angle of attack range of around = 5 upwards on the suction side. This corresponds to

46

6. R ESULTS OF O PTIMIZATION

0.25

0.13
0.45

0.12

0.4

0.47

0.13

0.1
0.15

0.3

0.2

0.35

0.12

0.05

0.11

0.05
0.1

0.11
0.3 0.3

0.25

0.1

0.45

0.3

0.09

0.45

0.4

0.1

0.1
0.15
0.2

0.1 0.15
0.2

0
0.2

0.25

0.09

0.35

0.08

0.08

0.15

0.07

0.3

0.35

0.06 0.05

0.06

0.1

0.35
0.4

0.4

0.3
0.25

0.04
3

4.5

5.5

6.5

0.04
3

3.5

4.5

5.5

6.5

10

15

20

(b) C P (,) soiled case.


2
1.75
1.5
1.25
1
0.75
0.5
Cl 0.25
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
20

N=9 free transition


N=0 forced transition at x/c = 0.1

0.04

(a) C P (,) clean case.

0.02

0.2

0.05

0.35

3.5

0.25

0.15 0.2 0.25

2
1.75
1.5
1.25
1
0.75
0.5
Cl 0.25
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
0

0.1

0.3

0.25

0.4

0.07

0.05

0.15
0.25

0.06

0.08

0.1

Cd

N=9 free transition


N=0 forced transition at x/c = 0.1

15

10

( )

(c) C l C d

(d) C l

Figure 6.8: Performance simulation results for the RK1-32 airfoil from the inviscid panel model.

RK132 CP
0.6

0.5

0.4

0.3

0.2

Inviscid clean
Double wake clean
Inviscid soiled
Double wake soiled

0.1

2.5

3.5

4
TSR

4.5

5.5

Figure 6.9: C P T SR of RK1-32 airfoil for a solidity of 0.1.

> 5 on the top surface and < 5 on the bottom surface. In the rest of the domain transition does occur
earlier, so an increase in drag is expected, but there is no influence on separation. Looking at the domain
where there is influence on separation, the trend seen in the results from the inviscid panel model seems
correct. This could be a valuable fact for an objective function optimizing for soiled conditions in the future.

47

0.6

0.6

0.5

0.5

0.4

0.4

CP

CP

6.2. A DAPTED OPTIMIZATION STRATEGIES

0.3

0.2

0.3

0.2

A23
A26
A32
NACA 0018

0.1

0
2.5

3.5

A23
A26
A32
NACA 0018

0.1

0
2.5

3.5

(a) C P T SR of RK1 series compared to the NACA 0018,


computed using the inviscid panel model.

TSR

TSR

(b) C P T SR of RK1 series compared to the NACA 0018,


computed using the double wake panel model.

Figure 6.10: C P T SR of RK1 series for a solidity of 0.1.

1
RK123
RK126
RK132
NACA 0018

0.9

0.8

0.7

0.7

0.6

0.6
xtr [x/c]

xtr [x/c]

0.8

0.9

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

0
20

15

10

0
[]

10

15

0
20

20

(a) Movement of transition point over a range of on the


top side of the airfoil.

RK123
RK126
RK132
NACA 0018

15

10

0
[]

10

15

20

(b) Movement of transition point over a range of on the


bottom side of the airfoil.

Figure 6.11: Movement of transition point on airfoils.

6.2. A DAPTED OPTIMIZATION STRATEGIES


In this section more investigation will be done into optimization using different strategies then in the previous
result section. The lessons learned from this previous campaign will be implemented in these trials. Four
cases were done to get a complete image of the possibilities and implications of the choices of objective
function and soiling case. The final result will be discussed in detail.
First of all the objective function derivation was revisited to see if this could offer some possibilities for improvement. The original derivation of the aerodynamic objective function shown in section 3.1 did not include C d . To see the influence of just optimizing for the lift slope (C l ), optimization strategy 2 and 3 were
performed with the objective of maximizing the lift slope.
To compare these strategies, these runs were also performed with the normal objective function used for the
RK1 series, C l /C d . These strategies (4 and 5) were performed for a fully clean and fully soiled (N=0) case.
The soiled and clean cases were chosen to see the effect of the weighting between clean and soiled used in
the RK1 series. All strategies are summarized in table 6.1.
During these optimizations, the maximum thickness constraint was turned off to ensure as much freedom as
possible. A first observation that can be made is that the results for the clean case (strategies 2 and 4) (figure
6.12a and 6.12c) have a higher thickness compared to the results from the soiled case. This is because the

48

6. R ESULTS OF O PTIMIZATION
Table 6.1: Summary of optimization strategies with objectives and weights for soiled (w s ) and clean (w c ) case.

Strategy 1
Strategy 2
Strategy 3
Strategy 4
Strategy 5

0.2
0.15
0.1
y/c 0.05
0
0.05
0.1
0.15
0.2
0.1

Aerodynamic objective
C l /C d
C l
C l
C l /C d
C l /C d

t/cmax=0.35

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

1.1

Structural objective
I xx /t
I xx /t
I xx /t
I xx /t
I xx /t

0.2
0.15
0.1
y/c 0.05
0
0.05
0.1
0.15
0.2
0.1

wc
0.5
1
0
1
0

t/cmax=0.27

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

x/c

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

x/c
Cl

(c) Strategy 4: C clean case.


d

1.1

1.1

(b) Strategy 3: C l soiled case.

t/cmax=0.38

x/c

(a) Strategy 2: C l clean case.


0.2
0.15
0.1
y/c 0.05
0
0.05
0.1
0.15
0.2
0.1

ws
0.5
0
1
0
1

1.1

0.2
0.15
0.1
y/c 0.05
0
0.05
0.1
0.15
0.2
0.1

t/cmax=0.27

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

x/c
Cl

(d) Strategy 5: C soiled case.


d

Figure 6.12: Highest performing outcomes of optimizations using different optimization objective formulation.

turbulent behavior of these profiles is not taken into account. Turbulent flow separation will occur earlier for
thick profiles. The clean case polars for these airfoils show good performance but the soiled case polars do
not. For these airfoils any kind of surface soiling would negatively influence the turbine performance. This
makes them not interesting for the present study.
Optimization strategies 2 and 3, using the objective of only C l , both show a dip in the top side of the airfoil.
In the RFOIL simulation at high positive angles of attack the flow skips this dip in the airfoil, while keeping
the high tail loading of the airfoil, the situation can be seen in figure 6.13a. This behavior will probably create
less lift and more drag than predicted by RFOIL, since there will already be separation at around 0.4c. The
region of reversed flow will produce drag and distort the rest of the flow. Also the upper part of the tail shape,
which was also seen in the RK1-32 (section 6.1.4), will likely produce more drag than predicted by RFOIL.
Looking at the pressure distribution in figure 6.13b the dip region shows a high pressure zone.
The most promising result of these four strategies is from the optimization strategy 5, using a fully soiled case
C

and the Cl objective. This airfoil shows high performance in the soiled case for which it was optimized, but
d
also for the clean case the performance is very high. The airfoil even outperforms the RK1-23 and RK1-26
which are both thinner. The airfoil is called RK2-27 (second generation, 27% thick), a comparison with the
NACA 0018 and the RK2-27 is shown in figure 6.14, the coordinates can be found in appendix A.
There are differences between the RK2-27 and the traditionally used NACA 0018. The first thing is the increased maximum thickness in the front part of the airfoil. This provides more freedom for structural designers to design a lighter and cheaper blade. Another difference is the camber that the RK2-27 has. The NACA
0018 and a lot of other previously used airfoils had symmetric profiles. The RK2-27 shows that this is not the
way to go and that camber can increase performance. The objective functions are the cause of this because
they only look at the length and slope of the linear part of the C l curve. The balance of loading can be regulated using fixed pitch. The final big difference is the tail shape. The RK2-27 uses a loaded tail to get to the
performance, this also resulted in a thinner tail than the NACA 0018.

6.2. A DAPTED OPTIMIZATION STRATEGIES

49

0.8

0.6

normalized Cp

0.4

0.2

0.2

(a) Airflow over 35% thick airfoil at = 10 from RFOIL.


0.4

0.6

0.1

0.2

0.3

0.4

0.5
x/c

0.6

0.7

0.8

0.9

(b) Normalized pressure distribution over 35% thick airfoil


from RFOIL.

y/c

Figure 6.13: Steady performance of 35% thick airfoil.

0.25
0.2
0.15
0.1
0.05
0
0.05
0.1
0.15
0.2
0.25
0.1

RK227
NACA 0018

0.1

0.2

0.3

0.4

0.5
x/c

0.6

0.7

0.8

0.9

1.1

Figure 6.14: Geometry of RK2-27 airfoil compared to the NACA 0018.

The performance of the RK2-27 in both clean and soiled conditions are shown in figure 6.15. This performance is computed using the inviscid panel model. The clean performance is higher compared to the previous RK1 generation. However the main advantage of this airfoil is the performance in soiled conditions. With
power coefficients up to 0.45 the deterioration in performance is small compared to the previous results. The
double wake results are shown in figure 6.16 for a solidity of 0.1. The red line shows the performance of the
RK2-27 and the blue line the performance of the RK1-26 from the previous generation. The clean case shows
higher performance at low TSR and about the same at a TSR of 4. For the RK1-26 the TSR of 5 was not simulated due to time constraints, but it could be the case that the performance passes the RK2-27. However in
soiled conditions the difference is clear, for the design TSR of 4 the C P of the RK2-27 is more than 0.3 higher.
This huge difference shows that the optimization for fully soiled conditions pays off. Also the optimal TSR
for this airfoil doesnt change significantly from clean to soiled conditions whereas this was the case for the
RK1-26 and RK1-32.
To find out where this big difference comes from the normal and tangential force predictions of the RK2-27
and the RK1-26 are shown in figure 6.17 for the design TSR of 4. The design TSR in the soiled case is shows a
large difference between the two airfoils in C P . The normal force F n shows that the RK1-26 blade is stalling
around 90 azimuth, the F t shows a rapid decrease here because of increased drag and decreased lift. Stall
makes the airfoil shed vortices downstream which it encounters at around 225 azimuth. These vortices decimate the lift force and thus the tangential force while the drag increases. This results in significant negative
values of F t . All this behavior is not observed with the RK2-27, deep stall in the upwind region is avoided,
reducing the vortex shedding compared to the RK1-26. In the downwind region the flow is therefore much
more stable and no losses are encountered and even positive torque can be produced here.

50

6. R ESULTS OF O PTIMIZATION

0.13

0.4

0.49 0.470.45

0.12

0.13

0.2
0.25

0.3
0.35

0.51

0.45

0.12

0.53

0.11

0.49

0.45
0.47

0.09

0.45
0.45

0.49

0.35
0.4

0.06

0.4

0.47

0.05

0.45

0.04
3

0.07

0.47

0.49

0.3

0.35

0.45

0.4

0.4

0.08

0.4

0.35

3.5

0.4

0.05

0.47
0.45

0.45

4.5

5.5

6.5

0.25
0.2

0.04
3

3.5

0.4

0.35

0.3

4.5

0.04

0.06

5.5

6.5

10

15

20

(b) C P (,) soiled case.


2
1.75
1.5
1.25
1
0.75
0.5
Cl 0.25
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
20

N=9 free transition


N=0 forced transition at x/c = 0.1

0.02

(a) C P (,) clean case.


2
1.75
1.5
1.25
1
0.75
0.5
Cl 0.25
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
0

0.15
0.25 0.2

0.35

0.4

0.51

0.45

0.06

0.05
0.1

0.3

0.47

0.07

0.2

0.45

0.1

0.35

0.51

0.08

0.25

0.11

0.1

0.49

0.35

0.3

0.4

0.09

0.4

0.08

0.1

Cd

N=9 free transition


N=0 forced transition at x/c = 0.1

15

10

( )

(c) C l C d

(d) C l

Figure 6.15: Performance simulation results for the RK2-27 airfoil from the inviscid panel model.
CP comparison
0.6

0.5

0.5

0.4

0.4

CP

CP

CP comparison
0.6

0.3

0.2

0.3

0.2

0.1

0.1
RK126
RK227

2.5

3.5
TSR

(a) Clean case

4.5

RK126
RK227
5

2.5

3.5
TSR

4.5

(b) Soiled case

Figure 6.16: Performance of RK2-27 compared to the performance RK1-26, calculated using the double wake model for a solidity of 0.1.

A critical note on the potential use of the transition point for roughness sensitivity estimation has to be made
here. The RK2-27 does not follow the statement made in the previous section, linking the movement to the
transition point to the robustness of the airfoil against soiling. Figure 6.18 shows the movement of the transition point in the clean case of the RK2-27 in red. The other airfoils from the RK1 serie and the NACA 0018 are
plotted in grey. As can be seen the theory suggested in the previous section does not hold for the RK2-27. This

6.2. A DAPTED OPTIMIZATION STRATEGIES

51

Normal Force for TSR = 4 and N = 0


1.5

Tangential Force for TSR = 4 and N = 0


RK227
RK126

RK227
RK126

0.25
0.2

1
Tangential Force []

Normal Force []

0.15

0.5

0.1
0.05
0
0.05
0.1

0.5

0.15
0.2
0

45

90

135

180
225
Azimuth []

270

315

360

45

90

135

(a) F N

180
225
Azimuth []

270

315

360

(b) F T

Figure 6.17: Normal and tangential forces of RK2-27 compared to the RK1-26, computed using the double wake model for TSR= 4.

means that this theory is not valid, at least not in all cases, it should be investigated further to prove valuable.
1

1
RK123
RK126
RK132
NACA 0018
RK227

0.9

0.8

0.7

0.7

0.6

0.6
xtr [x/c]

xtr [x/c]

0.8

0.9

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

0
20

15

10

0
[]

10

15

20

(a) Movement of transition point over a range of on the


top side of the airfoil.

RK123
RK126
RK132
NACA 0018
RK227

0
20

15

10

0
[]

10

15

20

(b) Movement of transition point over a range of on the


bottom side of the airfoil.

Figure 6.18: Movement of transition point on airfoils and the RK2-27.

7
F INAL R EMARKS
7.1. C ONCLUSIONS
This thesis focused on optimizing airfoils for VAWT and designing an efficient optimization procedure which
can be used for future optimizations for VAWT. Recalling the research goals:
1. To design an airfoil shape or family of shapes that is suitable for VAWT. This will give a general idea on
how VAWT airfoil should look and what characteristics are advised.
2. To validate the objective functions and optimization algorithm as an efficient optimization method to
design airfoils in future projects. For instance for different Reynolds numbers or different operating
conditions of the turbine.
The optimization strategy developed in this thesis demonstrates that VAWT airfoils can be designed with
aerodynamic performance in both clean and soiled conditions similar to its predecessors like the NACA 0018.
More importantly, the structural performance of the optimized airfoils is significantly higher.
The literature review showed a need of an optimization strategy for VAWT airfoils, genetic algorithms are
suitable to find optimal airfoils but require large computational power. To make genetic optimization feasible,
the optimal characteristics of a VAWT airfoil were derived. By considering the airfoil level the computational
power is reduced significantly with respect to full VAWT simulations, making genetic algorithms feasible.
The combination between the aerodynamic objective function and structural objective function generate
high performing airfoils for VAWT. Large thickness up to 27% (t/c) give structural designers more freedom and
possibilities then the traditionally used NACA 0018. The material needed for blade construction will decrease
which will lower investment costs for VAWT. The aerodynamic performance is not negatively influenced by
this increased thickness with C P s for clean conditions exceeding 0.5.
Blade soiling has a negative influence on power generation of a VAWT. Depending on the severity of roughness
the power output will drop significantly. VAWT designers should take this into account while developing
turbines. To make the airfoil robust to soiling the airfoil should be fully designed for these conditions. In this
thesis, soiled conditions are simulated in RFOIL using the critical amplification factor N equal to zero. This
modeling approach best mimics the effect of roughness on the airfoil.
This optimization process provides a quick airfoil design and performance evaluation procedure. There are
however some limits; due to the use of RFOIL as an airfoil analysis tool during the optimization no dynamic
effects can be simulated and assessed in the optimization. Next to this, there are shapes that are not accurately simulated by RFOIL. These shapes have either excessive curvature or are too thick. The limits have to
be taken into account when using the optimization process. Constraints can be set to avoid these designs but
have to be used with caution, not to eliminate feasible designs.
The full VAWT analysis models have limits. Especially the inviscid panel model does not perform well when
stall occurs due to high angles of attack or soiling. The double wake model and CFD simulations are taking
care of this phenomenon and agree well. Since the double wake model is computationally cheaper it is best
suited to do the simulations with.
53

54

7. F INAL R EMARKS

7.2. R ECOMMENDATIONS
Troposkien shaped VAWT have root airfoils close to the tower, these have different requirements than
the equator airfoils. This part of the blade could benefit from having blunt trailing edges which have
increased stiffness. It should be investigated how these flatback airfoils perform in RFOIL.
A good addition to the optimizer constraint definition is the possibility of setting boundaries not from
a reference population, but by manually setting boundaries. Currently the process of setting these
boundaries is highly iterative. If shapes for boundaries could be inserted manually this would save a lot
of time and make the boundaries more manageable.
The structural objective function is now optimizing for geometrical stiffness of the airfoil shape. In real
life the stiffness of a blade is determined for the most part by some load carrying structure inside the
blade. Examples of this are I-beams and box structures. It would be worthwhile to investigate if this can
be incorporated into the objective function. In this research a start was made on this topic, but was not
pursued further.
Research has to be done on how improve the simulation of roughness on airfoils. This can be done
either by doing experiments and validating the methods used in this thesis or by adding additional tools
to RFOIL to simulate roughness. An interesting topic for this could be the blowing/suction addition
that has already been made in RFOIL. If this can be validated as a roughness simulation tool it can be a
valuable tool. It has the added benefit that the boundary layer thickness can be increased, which also
happens when actual roughness is present.
The aerodynamic pitching moment generated by the airfoil can be taken into account in further studies on airfoil design. The aft loading of almost all airfoils from the optimization will amount to larger
pitching moments. This puts extra requirements on the structural design of the airfoil. It could be
incorporated into the aerodynamic optimization function as a constraint.
The movement of the transition point gives an indication of the sensitivity to soiling of the airfoil. This
can be implemented into an objective function to get airfoils which are less sensitive to roughness.
However this phenomenon should first be checked and investigated further since the principle did not
hold for all airfoils examined in this thesis.
The optimization strategy developed in this thesis should only be used if blade soiling is an issue. For
offshore HAWT the main concern is corrosion due to rain or hail impact on the blade. The lower TSR
of the VAWT combined with the impact direction on the blades make the VAWT less susceptible to this
type of corrosion. Before designing an airfoil for these conditions it should be investigated if surface
roughness is an issue.

B IBLIOGRAPHY
[1] A. Ghaderi, Image taken from:
windmills-of-nashtifan-in-iran/.

http://payvand.com/blog/blog/2010/11/21/photos-centuries-old-

[2] Image taken from: http://www.solar-and-wind-power-source.com/vertical-axis-wind-turbines.html.


[3] Image adapted from: http://www.ecosources.info/en/topics/Darrieus_vertical_axis_wind_turbine.
[4] Image taken from: http://en.wikipedia.org/wiki/Vertical_axis_wind_turbine.
[5] Quiet revolution, http://www.quietrevolution.com/index.htm.
[6] Image taken from: http://protorit.blogspot.nl/2011/06/wind-turbines-types-and-components.html.
[7] Scopus, Scopus - Analyze search results, www.scopus.com (2014).
[8] M. Snyder and N. Furukawa, Comparison of performance of Darrieus wind turbines having 12% and 21%
thick sections, Tech. Rep. 6 (Wichita State University, 1979).
[9] E. Kadlec, Characteristics of Future Vertical-Axis Wind Turbines, Tech. Rep. SAND79-1068 (Sandia National Laboratories, New Mexico, 1978).
[10] J. Healy, The influence of blade camber on the output of vertical-axis wind turbines, Wind Engineering
Vol.2, No.3, 146 (1978).
[11] J. Healy, The influence of blade thickness on the output of vertical axis wind turbines, Wind Engineering
Vol.2, No.1, 1 (1978).
[12] Y. Kato, K. Seki, and Y. Shimizu, Vertical axis wind turbine designed aerodynamically at tokai university,
(1980).
[13] R. Sheldahl and P. Klimas, Aerodynamic characteristics of seven symmetrical airfoil sections through 180degree angle of attack for use in aerodynamic analysis of vertical axis wind, Tech. Rep. SAND80-2114
(Sandia National Laboratories, New Mexico, 1981).
[14] P. Migliore and J. Fritschen, Comparison of NACA 6-series and 4-digit airfoils for Darrieus wind turbines,
Journal of Energy Vol. 7, No.4 (1983), 10.2514/3.48083.
[15] P. Klimas, Tailored airfoils for vertical axis wind turbines, Tech. Rep. SAND84-1062 (Sandia National Laboratories, New Mexico, 1984).
[16] A. Zervos, Aerodynamic Evaluation of Blade Profiles for Vertical axis wind turbine, in European Community Wind Energy Conference (1989) pp. 611616.
[17] D. Berg, Customized airfoils and their impact on VAWT cost of energy, in Windpower (1990).
[18] T. Ashwill, Measured data for the Sandia 34-meter vertical axis wind turbine, Tech. Rep. SAND91-2228
(Sandia National Laboratories, New Mexico, 1992).
[19] R. Galbraith, F. Coton, and J. Dachun, Aerodynamic design of Vertical Axis Wind Turbines, Tech. Rep. GU.
Aero. Rpt. No 9246 (Glasgow University, Glasgow, 1992).
[20] B. Kirke, Evaluation of self-starting vertical axis wind turbines for stand-alone applications, Ph.D. thesis,
Griffith University (1998).
[21] J. Vassberg, A. Gopinath, and A. Jameson, Revisiting the Vertical-Axis Wind-Turbine Design using Advanced Computational Fluid Dynamics, 43rd AIAA Aerospace Sciences Meeting and Exhibit (2005).
55

56

B IBLIOGRAPHY

[22] M. Claessens, The design and testing of airfoils for application in small vertical axis wind turbines, Masters thesis, Delft University of Technology (2006).
[23] R. Bourguet, G. Martinat, G. Harran, and M. Braza, Aerodynamic multi-criteria shape optimization of
VAWT blade profile by viscous approach, in Wind Energy, Proceedings of the Euromech Colloquium (2007).
[24] M. Islam, D. Ting, and A. Fartaj, Design of a special-purpose airfoil for smaller-capacity straight-bladed
VAWT, Wind Engineering Vol. 31, No. 6, 401 (2007).
[25] T. Carrigan, Aerodynamic shape optimization of a vertical axis wind turbine, Masters thesis, University
of Texas at Arlington (2010).
[26] M. Castelli and E. Benini, Effect of blade thickness on Darrieus Vertical-Axis Wind turbine performance,
CSSim 2011, 2nd International Conference on Computer Modelling and Simulation (2011).
[27] H. Sutherland, A retrospective of VAWT technology, Tech. Rep. SAND2012-0304 (Sandia National Laboratories, New Mexico, 2012).
[28] L. Danao, N. Qin, and R. Howell, A numerical study of blade thickness and camber effects on vertical
axis wind turbines, Proceedings of the Institution of Mechanical Engineers, Part A: Journal of Power and
Energy 226, 867 (2012).
[29] T. Carrigan, B. Dennis, Z. Han, and B. Wang, Aerodynamic Shape Optimization of a Vertical-Axis Wind
Turbine Using Differential Evolution, ISRN Renewable Energy Vol. 2012, 1 (2012).
[30] C. Simo Ferreira and B. Geurts, Aerofoil optimization for vertical axis wind turbines, Wind Energy
(2014), 10.1002/we.1762.
[31] G. de Oliveira, Wind Turbine Airfoils with Boundary Layer Suction , A Novel Design Approach, Masters
thesis, Delft University of Technology (2011).
[32] C. Simao Ferreira, M. Barone, A. Zanon, R. Kemp, and P. Giannattasio, Airfoil optimization for stall regulated vertical axis wind turbines, in AiAA Scitech, January (2015) pp. 116.
[33] T. Fukuda, New Airfoil Section for a Darrieus Wind Turbine, Masters thesis, Wichita State University
(1977).
[34] R. Eppler and D. Somers, A Computer Program for the Design and Analysis of Low-Speed Airfoils, Tech.
Rep. NASA TM-80210 (NASA Langley Research Center, Hampton, Virginia, 1980).
[35] D. Somers, Website Airfoils Incorporated, http://www.airfoils.com/index.htm (2014).
[36] R. Ramsay and G. Gregorek, Effects of Grit Roughness and Pitch Oscillations on the S824 Airfoil, Tech. Rep.
(Ohio State University, 1998).
[37] FloWind Corporation, Final Project Report: High-energy Rotor Development , Test and Evaluation, Tech.
Rep. SAND96-2205 (Sandia National Laboratories, San Rafael, California, 1996).
[38] E. Jacobs, Airfoil Section Characteristics as Affected by Variations of the Reynolds Number, Tech. Rep. No.
586 (NASA, 1937).
[39] F. Riegels, Aerofoil Sections, edited by D. Randall (Butterworths, London, 1961).
[40] Y. Kato, K. Seki, and Y. Shimizu, Vertical axis type wind power turbine, (1981).
[41] J. Strickland, The Darrieus Turbine: A Performance Prediction Model Using Multiple Streamtubes, Tech.
Rep. SAND75-0431 (Sandia National Laboratories, New Mexico, 1975).
[42] P. Migliore, W. Wolfe, and J. Fanucci, Flow curvature effects on Darrieus turbine blade aerodynamics,
Journal of Energy Vol.4, No.2, 49 (1980).
[43] P. Migliore and W. Wolfe, Some Effects of Flow Curvature on the Performance of Darrieus Wind Turbines,
17th AIAA Aerospace Sciencies Meeting (1979), 10.2514/6.1979-112.

B IBLIOGRAPHY

57

[44] C. Hirsch and A. Mandal, Flow curvature effect on vertical axis Darrieus wind turbine having high chordradius ratio, in European Wind Energy Conference, October (1984) pp. 405410.
[45] G. van Bussel, H. Polinder, and H. Sidler, TURBY: concept and realisation of a small VAWT for the built
environment, in EAWE/EWEA: The Science of making Torque from Wind, April (2004) pp. 509516.
[46] C. Simo Ferreira, The near wake of the VAWT: 2D and 3D views of the VAWT aerodynamics, Ph.D. thesis,
Delft University of Technology (2009).
[47] D. Ragni, C. Simo Ferreira, and M. Barone, Experimental and numerical investigation of an optimized
airfoil for vertical axis wind turbines, in AIAA Scitech: 32nd ASME Wind Energy Symposium, January
(2014) pp. 110.
[48] M. Islam, D. Ting, and A. Fartaj, Desirable airfoil features for smaller-capacity straight-bladed VAWT,
Wind Engineering Vol. 31, No. 6, 165 (2007).
[49] Gwind, Gwind, http://www.gwind.no/.
[50] Deepwind, The Deepwind Project, http://www.deepwind.eu/.
[51] INFLOW, Industrialization Setup of a Floating Offshore Wind Turbine, http://www.inflow-fp7.eu/.
[52] Nenuphar, Nenuphar Offshore Wind Turbines, http://www.nenuphar-wind.com/en.
[53] Spinfloat, Spinfloat, unleashing the potential of offshore power, http://www.spinfloat.com/en/.
[54] VertAx, Vertax Wind Ltd: Vertical Axis Wind Turbines, http://vertaxwind.com/.
[55] M. Drela and H. Youngren, XFOIL manual, http://web.mit.edu/aeroutil_v1.0/xfoil_doc.txt (2001).
[56] Image taken from: http://www.bladecleaning.com/problematica.htm.
[57] B. Kulfan and R. Hilton, A Universal Parametric Geometry Representation Method - CST, Journal of Aircraft Vol. 45, No. 1, 1 (2008).
[58] P. Stone, Image taken from: http://www.peterstone.name/Maplepgs/polyappr.html.
[59] Image taken from: http://commons.wikimedia.org/wiki/File:Front_pareto.svg.
[60] M. Drela, XFOIL: An analysis and design system for low Reynolds number airfoils, in Low Reynolds Number Airfoil Aerodynamics (1989).
[61] W. A. Timmer and R. P. J. O. M. van Rooij, Summary of the Delft University Wind Turbine Dedicated Airfoils, Journal of Solar Energy Engineering Vol. 125, 488 (2003).
[62] I. Paraschivoiu, Wind Turbine Design: With Emphasis on Darrieus Concept, 1st ed. (Polytechnic International Press, Canada, 2002).
[63] W. Timmer and A. Schaffarczyk, The effect of roughness at high Reynolds numbers on the performance of
aerofoil DU 97W300Mod, Wind Energy Vol. 7, No. 4, 295 (2004).
[64] W. Timmer and R. P. J. O. M. van Rooij, Thick airfoils for HAWTs, Journal of Wind Engineering and Industrial Aerodynamics Vol. 39, 151 (1992).
[65] C. van Dam, Conversation on airfoil transition, Conversation (2014).
[66] A. Zanon, P. Giannattasio, and C. S. a. Ferreira, A vortex panel model for the simulation of the wake flow
past a vertical axis wind turbine in dynamic stall, Wind Energy Vol. 16, No. 5, 661 (2013).
[67] A. Zanon, A vortex panel method for vawt in dynamic stall, Ph.D. thesis, University of Udine (2011).
[68] V. A. Riziotis and S. G. Voutsinas, Dynamic stall modelling on airfoils based on strong viscous-inviscid
interaction coupling, International Journal for Numerical Methods in Fluids Vol. 56, 185 (2008).

A
C OORDINATES OF R ESULTS

59

60

A. C OORDINATES OF R ESULTS

Table A.1: Upper side of RK1-23

x/c
0.00000
0.00122
0.00487
0.01093
0.01937
0.03015
0.04323
0.05853
0.07598
0.09549
0.11698
0.14033
0.16543
0.19217
0.22040
0.25000
0.28081
0.31270
0.34549
0.37904
0.41318
0.44774
0.48255
0.51745
0.55226
0.58682
0.62096
0.65451
0.68730
0.71919
0.75000
0.77960
0.80783
0.83457
0.85967
0.88302
0.90451
0.92402
0.94147
0.95677
0.96985
0.98063
0.98907
0.99513
0.99878
1.00000

y/c
0.00000
0.01085
0.02169
0.03251
0.04325
0.05380
0.06402
0.07376
0.08283
0.09107
0.09834
0.10451
0.10952
0.11330
0.11582
0.11710
0.11715
0.11602
0.11378
0.11053
0.10639
0.10152
0.09604
0.09011
0.08385
0.07736
0.07071
0.06397
0.05721
0.05048
0.04387
0.03749
0.03147
0.02590
0.02092
0.01658
0.01291
0.00992
0.00755
0.00572
0.00435
0.00334
0.00263
0.00216
0.00189
0.00181

Table A.2: Lower side of RK1-23

x/c
0.00000
0.00122
0.00487
0.01093
0.01937
0.03015
0.04323
0.05853
0.07598
0.09549
0.11698
0.14033
0.16543
0.19217
0.22040
0.25000
0.28081
0.31270
0.34549
0.37904
0.41318
0.44774
0.48255
0.51745
0.55226
0.58682
0.62096
0.65451
0.68730
0.71919
0.75000
0.77960
0.80783
0.83457
0.85967
0.88302
0.90451
0.92402
0.94147
0.95677
0.96985
0.98063
0.98907
0.99513
0.99878
1.00000

y/c
0.00000
-0.01276
-0.02532
-0.03749
-0.04912
-0.06007
-0.07027
-0.07963
-0.08808
-0.09556
-0.10199
-0.10726
-0.11128
-0.11395
-0.11521
-0.11503
-0.11344
-0.11055
-0.10648
-0.10143
-0.09559
-0.08918
-0.08238
-0.07533
-0.06816
-0.06094
-0.05373
-0.04659
-0.03958
-0.03277
-0.02625
-0.02015
-0.01459
-0.00970
-0.00560
-0.00237
-0.00005
0.00140
0.00204
0.00201
0.00148
0.00066
-0.00025
-0.00106
-0.00161
-0.00181

61

Table A.3: Upper side of RK1-26

x/c
0.00000
0.00122
0.00487
0.01093
0.01937
0.03015
0.04323
0.05853
0.07598
0.09549
0.11698
0.14033
0.16543
0.19217
0.22040
0.25000
0.28081
0.31270
0.34549
0.37904
0.41318
0.44774
0.48255
0.51745
0.55226
0.58682
0.62096
0.65451
0.68730
0.71919
0.75000
0.77960
0.80783
0.83457
0.85967
0.88302
0.90451
0.92402
0.94147
0.95677
0.96985
0.98063
0.98907
0.99513
0.99878
1.00000

y/c
0.00000
0.01133
0.02266
0.03398
0.04526
0.05648
0.06759
0.07853
0.08921
0.09952
0.10931
0.11843
0.12667
0.13384
0.13972
0.14411
0.14685
0.14781
0.14694
0.14427
0.13991
0.13402
0.12686
0.11871
0.10988
0.10069
0.09141
0.08231
0.07360
0.06544
0.05793
0.05111
0.04498
0.03948
0.03452
0.03000
0.02581
0.02186
0.01810
0.01453
0.01119
0.00818
0.00563
0.00368
0.00245
0.00203

Table A.4: Lower side of RK1-26

x/c
0.00000
0.00122
0.00487
0.01093
0.01937
0.03015
0.04323
0.05853
0.07598
0.09549
0.11698
0.14033
0.16543
0.19217
0.22040
0.25000
0.28081
0.31270
0.34549
0.37904
0.41318
0.44774
0.48255
0.51745
0.55226
0.58682
0.62096
0.65451
0.68730
0.71919
0.75000
0.77960
0.80783
0.83457
0.85967
0.88302
0.90451
0.92402
0.94147
0.95677
0.96985
0.98063
0.98907
0.99513
0.99878
1.00000

y/c
0.00000
-0.00847
-0.01713
-0.02616
-0.03561
-0.04546
-0.05557
-0.06569
-0.07551
-0.08469
-0.09288
-0.09977
-0.10511
-0.10873
-0.11051
-0.11045
-0.10858
-0.10500
-0.09984
-0.09329
-0.08557
-0.07693
-0.06767
-0.05809
-0.04850
-0.03922
-0.03053
-0.02265
-0.01574
-0.00992
-0.00519
-0.00151
0.00118
0.00302
0.00412
0.00460
0.00457
0.00415
0.00342
0.00248
0.00143
0.00037
-0.00059
-0.00136
-0.00186
-0.00203

62

A. C OORDINATES OF R ESULTS

Table A.5: Upper side of RK1-32

x/c
0.00000
0.00122
0.00487
0.01093
0.01937
0.03015
0.04323
0.05853
0.07598
0.09549
0.11698
0.14033
0.16543
0.19217
0.22040
0.25000
0.28081
0.31270
0.34549
0.37904
0.41318
0.44774
0.48255
0.51745
0.55226
0.58682
0.62096
0.65451
0.68730
0.71919
0.75000
0.77960
0.80783
0.83457
0.85967
0.88302
0.90451
0.92402
0.94147
0.95677
0.96985
0.98063
0.98907
0.99513
0.99878
1.00000

y/c
0.00000
0.01572
0.03132
0.04668
0.06169
0.07624
0.09022
0.10352
0.11601
0.12756
0.13800
0.14713
0.15474
0.16063
0.16460
0.16653
0.16636
0.16416
0.16010
0.15449
0.14772
0.14024
0.13250
0.12494
0.11790
0.11161
0.10619
0.10163
0.09779
0.09448
0.09143
0.08834
0.08493
0.08092
0.07608
0.07027
0.06342
0.05559
0.04697
0.03789
0.02880
0.02024
0.01279
0.00700
0.00332
0.00206

Table A.6: Lower side of RK1-32

x/c
0.00000
0.00122
0.00487
0.01093
0.01937
0.03015
0.04323
0.05853
0.07598
0.09549
0.11698
0.14033
0.16543
0.19217
0.22040
0.25000
0.28081
0.31270
0.34549
0.37904
0.41318
0.44774
0.48255
0.51745
0.55226
0.58682
0.62096
0.65451
0.68730
0.71919
0.75000
0.77960
0.80783
0.83457
0.85967
0.88302
0.90451
0.92402
0.94147
0.95677
0.96985
0.98063
0.98907
0.99513
0.99878
1.00000

y/c
0.00000
-0.01764
-0.03510
-0.05219
-0.06872
-0.08449
-0.09928
-0.11288
-0.12505
-0.13556
-0.14419
-0.15074
-0.15511
-0.15723
-0.15713
-0.15494
-0.15086
-0.14514
-0.13809
-0.13002
-0.12125
-0.11209
-0.10283
-0.09371
-0.08496
-0.07674
-0.06917
-0.06228
-0.05603
-0.05032
-0.04498
-0.03986
-0.03482
-0.02978
-0.02475
-0.01983
-0.01521
-0.01109
-0.00767
-0.00507
-0.00333
-0.00235
-0.00196
-0.00192
-0.00201
-0.00206

63

Table A.7: Upper side of RK2-27

x/c
0.00000
0.00122
0.00487
0.01093
0.01937
0.03015
0.04323
0.05853
0.07598
0.09549
0.11698
0.14033
0.16543
0.19217
0.22040
0.25000
0.28081
0.31270
0.34549
0.37904
0.41318
0.44774
0.48255
0.51745
0.55226
0.58682
0.62096
0.65451
0.68730
0.71919
0.75000
0.77960
0.80783
0.83457
0.85967
0.88302
0.90451
0.92402
0.94147
0.95677
0.96985
0.98063
0.98907
0.99513
0.99878
1.00000

y/c
0.00000
0.01332
0.02660
0.03977
0.05278
0.06557
0.07803
0.09008
0.10156
0.11230
0.12209
0.13069
0.13784
0.14329
0.14684
0.14835
0.14777
0.14518
0.14075
0.13477
0.12761
0.11966
0.11134
0.10299
0.09490
0.08725
0.08014
0.07357
0.06750
0.06183
0.05647
0.05134
0.04638
0.04153
0.03678
0.03214
0.02761
0.02323
0.01903
0.01508
0.01146
0.00827
0.00563
0.00364
0.00240
0.00198

Table A.8: Lower side of RK2-27

x/c
0.00000
0.00122
0.00487
0.01093
0.01937
0.03015
0.04323
0.05853
0.07598
0.09549
0.11698
0.14033
0.16543
0.19217
0.22040
0.25000
0.28081
0.31270
0.34549
0.37904
0.41318
0.44774
0.48255
0.51745
0.55226
0.58682
0.62096
0.65451
0.68730
0.71919
0.75000
0.77960
0.80783
0.83457
0.85967
0.88302
0.90451
0.92402
0.94147
0.95677
0.96985
0.98063
0.98907
0.99513
0.99878
1.00000

y/c
0.00000
-0.01111
-0.02225
-0.03344
-0.04463
-0.05573
-0.06657
-0.07696
-0.08667
-0.09547
-0.10315
-0.10957
-0.11462
-0.11827
-0.12051
-0.12137
-0.12091
-0.11915
-0.11612
-0.11183
-0.10632
-0.09962
-0.09183
-0.08309
-0.07363
-0.06372
-0.05372
-0.04395
-0.03478
-0.02648
-0.01927
-0.01329
-0.00856
-0.00503
-0.00257
-0.00102
-0.00018
0.00012
0.00005
-0.00023
-0.00063
-0.00105
-0.00143
-0.00173
-0.00192
-0.00198

B
AIAA PAPER
The preliminary findings from this thesis were presented on the 33r d Wind Energy Symposium at the AIAA
SciTech conference in January 2015 in Kissimmee Florida. The full paper can be found in this appendix.

65

AIAA 2015-0722
AIAA SciTech
5-9 January 2015, Kissimmee, Florida
33rd Wind Energy Symposium

Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

Airfoil optimization for stall regulated vertical axis


wind turbines
Carlos Simao Ferreira

Matthew Barone

Delft University of Technology - DUWind

Sandia National Laboratories

Delft, 2629HS, The Netherlands

Albuquerque, NM 87185, USA

Alessandro Zanon

Rody Kemp

AIT Austrian Institute of Technology GmbH

Delft University of Technology - DUWind

Vienna, 1210, Austria

Delft, 2629HS, The Netherlands

Pietro Giannattasio
University of Udine - Dipartimento di Ingegneria Elettrica Gestionale e Meccanica
Udine, 33100, Italy

Multi-megawatt Vertical Axis Wind Turbines (VAWTs) have inherent design and operational advantages that make them relevant for floating oshore applications. Many oshore
VAWT concepts use stall for power regulation. Stall regulation imposes several constrains
in the aerodynamic, structural and generator design. Due to the azimuthally varying and
unsteady aerodynamics experienced by a VAWT, designing an airfoil for stall regulation is
still a significant challenge. In this work, a family of airfoils for stall regulated VAWT is defined through numerical airfoil optimization, based on the original work of Sim
ao Ferreira
and Geurts [20]; the optimization is multi-objective, optimizing structural and aerodynamic
performance. The control performance is not yet implemented in the function. The optimization is performed at Reynolds numbers representative of multi-megawatt VAWT. The
performance of the VAWT, including operation in dynamic stall, is evaluated with an unsteady double-wake viscous-inviscid panel method and CFD simulations. The performance
of the optimized airfoils is compared against a conventional airfoil, namely the NACA 0018
airfoil. The stall regulatory performance is assessed to provide insight for future optimizations. The aerodynamic performance is evaluated using three dierent numerical models:
a single wake panel model coupled to airfoil polar data; a viscous-inviscid double wake
panel model, capable of simulating dynamic stall; and an eulerian RANS CFD model. The
airfoils are also design for a robust performance in the case of surface roughness. The
preliminary results show that the design for surface roughness conflicts with the design for
dynamic stall control. An extensive study of multi-objective optimisations with dierent
weights of the dierent elements of aerodynamic performance is presented.

Nomenclature

Cd
Cf

Angle of attack
Range of angle of attack (representative of tip speed ratio)
Tip speed ratio
Bc
Solidity of turbine ( 2R
)
2D drag coefficient
Skin friction coefficient

Sandia National Laboratories is a multi-program laboratory managed and operated by Sandia Corporation, a wholly owned
subsidiary of Lockheed Martin Corporation, for the U.S. Department of Energys National Nuclear Security Administration
under contract DE-AC04-94AL85000.

1 of 16
American Institute of Aeronautics and Astronautics
Copyright 2015 by Carlos Simao Ferreira, Matthew Barone, Alessandro Zanon, Rody Kemp, Pietro Giannattasio. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.

Cl
CP
C l
Ncrit
xtr
B
c
N
R
s
t/c

2D lift coefficient
Power coefficient
Lift slope
Critical amplification factor
(Forced) transition point
Number of blades
Airfoil chord
Ncrit
Turbine radius
Outcome of objective function
Relative thickness of airfoil

Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

I.

Introduction

VAWTs have typically used symmetric airfoils, notably the NACA 4-series (in particular the NACA 0012,
NACA 0015 and NACA 0018, see Sheldahl and Klimas [17] and Timmer [23]). Examples of research (design,
numerical analysis and experimental test) on VAWT airfoils are the ones developed at Sandia National Laboratories (see Klimas [13]), at Glasgow University (see Galbraith et al. [9]) and the work at Delft University
of Technology for small scale VAWTs (see Claessens [3]). The research at Sandia National Laboratories and
Glasgow University achieved two milestones: the determination of desirable section characteristics for lowering the cost of energy [13]; and the coupling of the airfoil design optimisation with a vorticity based model of
the rotor [9], eliminating many of the incorrect assumptions of streamtube models. The work of Klimas [13],
citing the work of Sullivan [22], includes, as design drivers, the impact of stall regulation (modest values
of maximum lift coefficient with relatively sharp stall, [13]) and low drag (low zero lift drag coefficients,
and [...] wide drag buckets, Klimas [13]). The work of Galbraith et al. [9] integrates the improvements
in modelling the induction field developed by Jiang et al. [11] (see also Basuno et al. [2] and Coton et al.
[4]). Other design eorts by Migliore and Fritschen [15] and Zervos [30] aimed to include the flow curvature
phenomenon into the airfoil design for VAWT. In both methods, where original airfoil shapes were adapted
with some sort of curvature, the impact of the method heavily relied on the turbine solidity (c/R).
Several designs of VAWTs have in the past achieved maximum power coefficients in line with current
Horizontal Axis Wind Turbines (HAWTs) (see Ashwill [1], Maydew and Klimas [14] ). However, to decrease
the cost of energy of VAWTs it is necessary to design airfoils that are both aerodynamically and structurally
efficient. A way to improve structural efficiency is to introduce stall regulation into a turbine. Turbines with
constant rotational speed will reduce tip speed ratio ( ), increasing angles of attack, thus stalling the airfoil
and reducing loads and power. For a stall regulated VAWT, the airfoil must balance high performance at
rated wind speed with a sharp loss of efficiency with decreasing tip speed ratio. Airfoils with sharp drops in
Cl and rapid increasing Cd at stall are interesting for stall regulated VAWT. This behavior can induce extra
vibrations in the structure due to larger oscillating separation, this must be kept in mind but is outside the
scope of this paper.
Airfoil optimization for a wind turbine should account for possible early turbulent transition due to surface
soiling and/or high turbulence. Early transition, especially for thick airfoils, can lead to turbulent separation
and loss of aerodynamic efficiency by both increasing drag and decreasing lift; the works of Van Rooij and
Timmer [26] and Timmer and Schaarczyk [24] show the current lack of knowledge, the uncertainty of the
impact of this lack of knowledge and some ideas on how to model that impact. In HAWT airfoil design,
dierent approaches are used, ranging from optimising for turbulent flow to only considering the decrease
of maximum glide ratio and shift of angle of attack of maximum glide ratio (for control and uncertainty
quantification). The optimisation implemented in this work accounts for the possible impact of leading edge
forced transition, roughness and/or high turbulence.

II.

Methods

The approach for this research is composed of four steps:


1. Generate airfoils for VAWT following the methodology proposed by Simao Ferreira and Geurts [20],
using a dual-objective genetic optimisation: an aerodynamic optimisation function and a structural

2 of 16
American Institute of Aeronautics and Astronautics

optimisation function.
2. Calculate the power curve of a two-bladed 2D VAWT with the generated airfoils, using three dierent
simulation approaches.
(a) A potential-flow panel-model free single-wake coupled to airfoil viscous polars, where the impact
of flow separation in the induction and dynamic stall are neglected.
(b) A viscous-inviscid interaction double-wake unsteady panel model capable of modelling a VAWT
in dynamic stall (see [29]).
(c) A RANS CFD simulation model (SST-DES hybrid RANS/LES turbulence model).

Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

3. Compare the performance with an airfoil used very commonly in VAWTs, namely the NACA 0018.
A.

Airfoil generation

The airfoil shape is the result of an optimization for aerodynamic and structural performance (Simao Ferreira
and Geurts [20]). Departing from an initial population of airfoils, through a genetic algorithm, the shape
is synthesized and evaluated by two objective functions. The optimization algorithm used in this work
was developed by de Oliveira [5] and Sim
ao Ferreira and Geurts [20]; it is constructed around the NSGA-II
algorithm as implemented in MATLAB R2011. The most relevant elements of the algorithm are described by
de Oliveira [5]. The structural objective function was defined as the bending stiness in flap-wise direction
of the airfoil per wall-thickness, in reference to the centroid of the airfoil. The aerodynamic function is
the one presented by Sim
ao Ferreira and Geurts [20]. A simplification is made relating to the range and
distribution of angles of attack over the turbine revolution. The range of angles of attack will vary for
Bc
dierent tip-speed ratios and dierent values of 2R
(assuming rigid, fixed pitch blades); in the present study,
a single evaluation interval of angles of attack ( = 15 ) is accounted for. This range of angles of attack is
representative of a VAWT at a of about 4. This range is tested against a range of possible angles of
attack 20 < < 20 . The core of the objective function is given by Equation 1, where b() is a weighting
function based on the frequency of occurrence of angles of attack over a rotation presented in Simao Ferreira
and Geurts [20].

saero =

00

BB
BB
max BB
@@

+R r

Cl d C
C
C , for
+R r
A
Cd b() d

20 <

< (20

C
C
)C
A

(1)

In Equation 1, makes sure the whole possible angle of attack range 20 < < 20 is covered. Cl
and Cd are the lift slope and drag coefficient as a function of angle of attack . These are calculated using
aerodynamic coefficients taken from RFOIL [25]. By doing this for a fixed Reynolds number of 5 106 , any
variation of dynamic pressure with the azimuthal position is not taken into account. This was possible to
the relatively lower impact shown in a secondary analysis, where it was found that the dynamic pressure
variations mostly average out across leeward and windward regions. To take into account the behavior of
turbulent transition due to soiling of the airfoil, Equation 1 is evaluated for a clean case and a soiled case.
The clean case being free transition and a critical amplification factor Ncrit of 9 and the soiled case being
forced transition at 10% of the airfoil (both sides) and Ncrit equal to 0. The factor Ncrit is a measure of
stability of the ambient flow, low Ncrit flows will transition to turbulent flow earlier than high Ncrit flows.
The final aerodynamic objective function value takes a weighted average of clean and turbulent performance,
given by Equation 2.
saero =

0.5
sclean

1
+

0.5
ssoiled

(2)

To avoid very thin trailing edges, a constraint is placed on the trailing edge angle. An optimization was
run both with and without this constraint. The minimum trailing edge angle was set at 8 .

3 of 16
American Institute of Aeronautics and Astronautics

B.

Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

1.

Simulation of the VAWT performance


Simulation with a potential-flow panel-model free single-wake coupled to airfoil viscous polars

The induction field for the 2D VAWT is calculated using the 2D unsteady panel method of [18]. The solutions
are calculated for several values of Cl .c/R, using a NACA 0015. This approximation is valid due to the fact
that blade sections with the same Cl .c will, for the same azimuthal distribution of perceived velocity, have
the same azimuthal distribution of circulation and shed the same wake, and thus the same induction field and
perceived velocity. When non-dimensionalising the length scale of the solution by R, the non-dimensional
induction field of a 2D VAWT will be the same for rotors with dierent blade sections if Cl .c/R is the same.
The solutions of induction over the rotation are used as the solution for the optimised viscous airfoils for
equal viscous Cl .c/R, calculated at the linear region of the lift slope; this approach is conservative in the
calculation of torque and power, as it penalizes the viscous airfoil with a higher induction as compared with
the inviscid case. The induction solution is used to determine the angle of attack and viscous loading over
the rotation, using the airfoils polar data. For the purpose of this analysis and range of c/R, it is acceptable
to ignore the impact of blade vortex interaction and other eects at the chord scale. This approach is
not however correct when dynamic stall is a relevant phenomena. Additionally, flow curvature eects are
approximated by using a correction for the added circulation.
2.

Simulation with a viscous-inviscid double wake panel model for separated flow simulation

The double-wake viscous-inviscid panel model is described in detail in [29] together with its experimental
validation.
The panel equations for the unsteady potential flow are solved together with the integral boundary
layer equations on the airfoil surface. The unsteady panel model has been extended to separated flows by
introducing a second panel originating at the separation point. This double-wake approach (see [12, 16, 27,
28]) is based on the concept that at separation, vorticity is continuously released in the flow along a strong
shear layer that bounds the separation bubble. The vorticity release is modelled as the shedding of vortex
blobs at the separation point, the location of which is predicted by the boundary layer model.
The boundary layer model uses Drelas integral approach for steady flows (see [6]). For the laminarturbulent transition, the model uses the simplified version of the en method developed by [8]. The potential
flow and boundary layer equations are not solved simultaneously, but a semi-inverse iterative scheme is
adopted.
The separation criterion used in the present model is based on the value of the skin friction coefficient,
Cf (flow separates when Cf becomes negative). This criterion has been shown to be reliable in previous
works ([16]), but it fails when applied to pitching airfoils at medium-low Reynolds numbers, because it does
not allow flow reattachment after a deep stall at the leading edge has occurred. An improved criterion is
implemented in the [29] to identify the condition under which the turbulent reattachment becomes possible.
The double-wake model was validated in [29] with reference to steady and pitching airfoils NACA 0015
and NACA 0012 and to a VAWT in dynamic stall, on the basis of numerical and experimental data provided
in [7, 10, 16, 19, 21, 29].
The double-wake model is able to capture the complex evolution of the vorticity field associated to the
energy conversion process of the VAWT, and the predicted values of the forces are comparable to the ones
provided by the much more expensive CFD approaches.
3.

Simulation with a RANS CFD model

Nalu is a low-Mach number computational fluid dynamics code that can be configured to solver the incompressible Navier-Stokes equations using various turbulence model options including RANS, LES, and hybrid
RANS-LES models. The code uses an unstructured mesh algorithm with a fully implicit solver designed
and demonstrated to efficiently scale on many-core high-performance computing clusters. Nalu allows for
non-conforming mesh interfaces, which has been exercised in the present simulations to create a circular rotating mesh block containing the VAWT rotor. In the current simulations, the SST-DES hybrid RANS/LES
turbulence model has been applied, without any explicit model for boundary layer transition. In this way, at
present, the CFD model provides a prediction for worst case boundary layer transition, where the boundary
layer is immediately turbulent very close to the leading edge.

4 of 16
American Institute of Aeronautics and Astronautics

III.

Results and discussion

Figures 1, 2 and 3 show the geometry, polar and performance simulation results (CP( , ) and CP( ) for
= 0.1) for the 3 airfoils that resulted from the optimization (A-23, A-26 and A-32, last two digits indicate
relative thickness). The performance simulation are based in the single wake panel model simulations. In
this section each airfoil will be analysed using the results of the double wake panel code and the inviscid
panel code. For the A-23 airfoil CFD results from NALU will complement the analysis. The tangential force
(non dimensional) is used as an indicator of performance.

Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

A-23
The A-23 is a 23% thick airfoil with a slight positive camber, see figure 1a. This airfoil resulted from the
optimization without the trailing edge constraint, which left it very thin. For this airfoil, the influence of
pitch was not investigated. The eect of roughness is shown in figure 4, for a of 4 the performance drops,
but only by a small amount. This decrease in performance can largely be contributed to the increase in drag
in turbulent flows, next to this some small separation is occuring at the trailing edge. For a of 3, the drop
in performance is more significant due to the deep stall the airfoil experiences. Figure 13 shows the vorticity
of a turbine with the A-23 airfoil at = 4 computed with the CFD model. Performance found from this
model show CP = 0.39 for = 4 and CP = 0.23 for = 3.
A-26
The A-26 is a 26% thick airfoil with positive camber. As can be seen in figure 2a, the trailing edge has
sufficient thickness due to the constraint placed on this. Looking at the Cl polar and keeping in mind the
(asymmetric) angle of attack distribution, it would be logical to add pitch to an airfoil to avoid stall. The
eect of 4 pitch is shown in figure 5, a shift in forces between upwind and downwind part of the rotor.
At a of 3, stall is avoided on the upwind part (figure 5a). Pitch of 4 also reduces peak loads on blade
as seen in figure 5d and 5c with the normal forces more equally divided between upwind and downwind. A
comparison between the clean performance of A-26 and NACA 0018 can be found in figure 6a. The impact
of dierent soiling cases is shown in figure 6b. Two of the soiling cases exhibit stall in the downwind region
reducing the power output drastically. In these cases the forward transition point leads to a boundary layer
with a reduced ability to cope with the adverse pressure gradient and deep stall occurs.
A-32
The A-32 is a 32% thick airfoil, also with positive camber. Its main feature is the bulging of the top surface
at the tail. This unexpected result was pursued for its fair performance prediction with the inviscid panel
model (figure 2c and 2d) and his huge potential benefits in the structural design. Again a pitch of 4 was
applied, but this didnt have a beneficial eect on the performance. Stall was present on the downwind part
of the rotation as can be seen in figure 7a. The clean performance of the A-32 is compared to the NACA
0018 in figure 8a. The eect of the dierent soiling cases is shown in figure 8b. The A-32 performance is
very sensitive to the soiling cases. With N=0 the airfoil stalls upwind and a negative CP is observed. The
least severe soiling case (N=4, xtr = 0.2) shows almost no dierence with the clean case.
Figure 9 shows the clean performance of the three airfoils discussed compared to the NACA 0018. The
performance for = 4, which is the design- , is around 4% higher for all optimized airfoils despite the
fact that all airfoils have higher thickness and are therefore structurally much more favorable. With the
traditionally used NACA 4 series airfoils, increasing thickness leads to decreasing aerodynamic performance,
with these optimized airfoils this is not the case. The simulations show good performance in clean case but
it decreases when simulating in soiled conditions, the severity in drop depends significantly on the choice of
N factor and forced transition point. For the worst case (N=0), the performance drops with thickness with
performance at = 4 of the A-23 of 0.4, the A-26 of 0.12 and a negative CP for the A-32. However for
the best case of soiling tested, the A-32 (N=4 and forced transition at 0.2 x/c) had a CP of 0.52, which
was higher than for instance the A-26 with 0.43. The simulation of roughness should be a subject of further
investigation.

5 of 16
American Institute of Aeronautics and Astronautics

Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

Comparison of Double-Wake and CFD Model Solutions


In this work we have employed two models that are capable of modeling the performance of VAWT rotors
into the dynamic stall regime the double-wake potential flow model, and the hybrid RANS-LES CFD model.
In figure 11, we compare the normal and tangential blade force predictions for these two methods for zero
blade pitch angle for the RK26 airfoil. Given that the method for boundary layer transition specification
is quite dierent between the two models, the results are remarkably similar. Good agreement between the
two simulation methods is obtained at both TSR=3 and TSR=4. The major dierence is that the CFD
results, unlike the double-wake results, exhibit strong high-frequency oscillations during the revolution; these
oscillations are associated with flow separation and subsequent vortex shedding that is observed during the
simulations, these are not present in the double wake simulation since the results are averaged over several
revolutions.
Comparison of results using the two methods for the RK32 airfoil is shown in Figure 12. Overall, the
agreement in force predictions is reasonable between the two approaches, although there are now some
noticeable dierences in the predictions as the blade travels upwind and stalls near theta=90 deg. Again,
more dramatic force oscillations associated with separation are observed in the CFD results.

IV.

Conclusions

The preliminary results show that the optimized airfoils achieve a high power coefficient with thick airfoils
(t/c > 0.23) when assuming free transition. The strategy of setting a low N-factor in the transition model to
account for leading edge roughness/higher flow turbulence resulted in thick airfoils from the optimization.
However, when simulated in these soiled conditions, these airfoils still show significant performance losses.
Further research will focus on how to better incorporate the eect of blade soiling into the optimization
and analysis methods.
The preliminary results appear to indicate that the main challenges for the design of stall regulated airfoils
are: 1) uncertainty of the surface soiling lifetime and how to model it (as also seen for HAWTs); 2) near
wake development; and 3) the eect of blade-vortex interaction in transition. Despite these limitations, the
application of the new optimisation objective function delivers airfoils with higher aerodynamic performance
and higher structural performance which are still suitable for stall regulation. The performance of the airfoils
is verified by three independent numerical simulation codes.

V.

Acknowledgments

This work is made possible in part due to funds from Technology Foundation STW - Veni grant: Innovational Research Incentives Scheme Experimental and numerical multi-scale investigation of VAWT rotor-wake
systems for oshore applications, the FP7-Deepwind project and the U.S. Department of Energy project:
Innovative Oshore Vertical-Axis Wind Turbine.

Ashwill, T. (1992). Measured data for the sandia 34-meter vertical axis wind turbine. Technical Report
SAND91-2228, Sandia National Laboratories.

Basuno, B., Coton, F., and Galbraith, R. (1992). A prescribed wake aerodynamic model for vertical axis
wind turbines. Proc. Institution Mechanical Engineers, Part A: Journal of Power and Energy, 206:159166.

Claessens, M. (2006). The design and testing of airfoils for application in small vertical axis wind turbines.
Masters thesis, Faculty of Aerospace Engineering, Delft University of Technology.

Coton, F., Jiang, D., and Galbraith, R. (1994). Unsteady prescribed wake model for vertical axis wind
turbines. Proc. Institution Mechanical Engineers, Part A: Journal of Power and Energy, 208:1316.

de Oliveira, G. (2011). Wind turbine airfoils with boundary layer suction, a novel design approach. Masters
thesis, Delft University of Technology.

6 of 16
American Institute of Aeronautics and Astronautics

Drela, M. (1989). Xfoil: an analysis and design system for low Reynolds number airfoils. In Conference
Proceedings on Low Reynolds Number Aerodynamics. University of Notre Dame, Indiana.

Drela, M. (XFOIL 6.9). http://web.mit.edu/drela/public/web/xfoil/.

Drela, M. and Giles, M. (1987). Viscous-inviscid analysis of transonic and low Reynolds number airfoils.
AIAA Journal, 25:13471355. DOI: 10.2514/3.9789.

Galbraith, R., Coton, F., and Dachun, J. (1992a). Aerodynamic design of vertical axis wind turbines.
Technical Report GU Aero 9246, University of Glasgow.

10

Galbraith, R., Gracey, M., and Leitch, E. (1992b). Summary of pressure data for thirteen aerofoils on
the University of Glasgow aerofoil database. G.U. Aero Report 9221.

Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

11

Jiang, D., F.N.Coton, and Galbraith, R. M. (1991). A fixed vortex model for vertical axis wind turbines
including unsteady aerodynamics. Wind Engineering, 15(6):348360.

12

Katz, J. (1981). A discrete vortex method for the non-steady separated flow over an airfoil. J. of Fluids
Mechanics, 102:315328. DOI: 10.1017/S0022112081002668.

13

Klimas, P. C. (1984). Tailored airfoils for vertical axis wind turbines. Technical Report SAND84-1062,
Sandia National Laboratories and United States. Dept. of Energy.

14

Maydew, R. C. and Klimas, P. (1981). Aerodynamic performance of vertical and horizontal axis wind
turbines. Journal of Energy, 5(3):189190.

15

Migliore, P. and Fritschen, J. (1983). Comparison of NACA 6-series and 4-digit airfoils for Darrieus wind
turbines. Journal of Energy, 7.

16

Riziotis, V. and Voutsinas, S. (2008). Dynamic stall modelling on airfoils based on strong viscous inviscid interaction coupling. International Journal for Numerical Methods in Fluids, 56:185208. DOI:
10.1002/fld.1525.

17

Sheldahl, R. E. and Klimas, P. C. (1981). Aerodynamic characteristics of seven symmetrical airfoil sections
through 180-degree angle of attack for use in aerodynamic analysis of vertical axis wind turbines. Technical
Report SAND81-2114, Sandia National Laboratories.

18

Sim
ao Ferreira, C. (2009). The near wake of the VAWT: 2D and 3D views of the vawt aerodynamics.
PhD thesis, Delft University of Technology. ISBN/EAN:978-90-76468-14-3.

19

Sim
ao Ferreira, C., Bijl, H., van Bussel, G., and van Kuik, G. (2007). Simulating dynamic stall in a 2D
VAWT: Modeling strategy, verification and validation with particle image velocimetry data. Journal of
Physics: Conference Series, 75(1):012023.

20

Sim
ao Ferreira, C. and Geurts, B. (2014). Aerofoil optimisation for the vertical axis wind turbine. Wind
Energy - online. online.

21

Sim
ao Ferreira, C., van Bussel, G., van Kuik, G., and Scarano, F. (2009). Visualization by PIV of
dynamic stall on a vertical axis wind turbine. Experiments in Fluids, 46(1):97108.

22

Sullivan, W. (1979). Economic analysis of darrieus vertical axis wind turbine systems for teh generation
of utility grid electric power - vol. 2: The economic optimization model. Technical report, Sandia National
Laboratories.

23

Timmer, W. (2008). Two-dimensional low-reynolds number wind tunnel results for airfoil naca 0018.
Wind Engineering, 32(6):525537.

24

Timmer, W. and Schaarczyk, A. (2004). The eect of roughness at high reynolds numbers on the
performance of aerofoil du 97-w-300mod. Wind Energy, 7(4):295307.

25

van Rooij, R. (1996). Modification of the boundary layer calculation in rfoil for improved airfoil stall
prediction. Technical report, Delft University of Technology.
7 of 16
American Institute of Aeronautics and Astronautics

26

Van Rooij, R. and Timmer, W. (2003). Roughness sensitivity considerations for thick rotor blade airfoils. Journal Of Solar Energy Engineering - Transactions-American Society Of Mechanical Engineers,
125(4):468478.

27

Vezza, M. and Galbraith, R. (1985). An inviscid model of unsteady aerofoil flow with fixed upper surface
separation. J. Num. Meth. in Fluids, 5(6):577592. DOI: 10.1002/fld.1650050607.

28

Voutsinas, S. and Riziotis, V. (1999). A viscous-inviscid interaction model for dynamic stall simulations
on airfoils. AIAA Paper 99-0038.

29

Zanon, A., Giannattasio, P., and Sim


ao Ferreira, C. (2013). A vortex panel model for the simulation
of the wake flow past a vertical axis wind turbine in dynamic stall. Wind Energy, 16:661680. DOI:
10.1002/we.1515.

Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

30

Zervos, A. (1989). Aerodynamic Evaluation of Blade Profiles for Vertical axis wind turbine. In EWEC,
pages 611616.

8 of 16
American Institute of Aeronautics and Astronautics

0.6
N=9 free transition
N=0 forced transition at x/c = 0.1

0.55
0.5
0.2
0.15
0.1
x/c 0.05
0
0.05
0.1
0.15
0.2
0.1

0.45
0.4

t/cmax=0.23

0.3

0.25
0.2
0

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

0.15

1.1

y/c
Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

0.35

0.1
0.05
0
2.5

3.5

4.5

5.5

6.5

(b) CP(

(a) Geometry.
0.13

0.35
0.49

0.12
0.11

0.47 0.45

0.4

0.3

0.35

0.09

0.47 0.45

0.49

0.06

0.4
0.4

0.25
0.4

0.06

0.35
0.4

0.45

0.05
0.25

0.3 0.35

0.4

4.5

5.5

6.5

0.04
3

0.2

0.3

3.5

0.35

4.5

(c) CP(

clean case.

, )

0.06

0.08

Cd

(e) Cl

5.5

6.5

10

15

20

(d) CP(

N=9 free transition


N=0 forced transition at x/c = 0.1

0.04

0.35

0.3

0.45

0.05

0.1
0.2 0.15

0.3

0.07 0.3 0.35

0.47
0.4

0.02

0.08

0.35
0.4

0.47
0.45

2
1.75
1.5
1.25
1
0.75
0.5
Cl 0.25
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
0

0.15

0.25

0.09

0.49

3.5

0.2

0.1

0.25
0.3

0.04
3

0.10.05

0.25

0.11
0.4

0.07

= 0.1.

0.35

0.4

0.12

0.51

0.1

0.08

for

0.13

0.1
0.15
0.2

0.25

0.3

0.1

2
1.75
1.5
1.25
1
0.75
0.5
C 0.25
l
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
20

, )

soiled case.

N=9 free transition


N=0 forced transition at x/c = 0.1

15

10

( )

Cd

(f) Cl

Figure 1: Geometry, polar and performance simulation results for the A-23 airfoil.

9 of 16
American Institute of Aeronautics and Astronautics

0.6
N=9 free transition
N=0 forced transition at x/c = 0.1

0.55
0.5
0.2
0.15
0.1
x/c 0.05
0
0.05
0.1
0.15
0.2
0.1

0.45
0.4

t/cmax=0.26

0.3

0.25
0.2
0

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

0.15

1.1

y/c
Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

0.35

0.1
0.05
0
2.5

3.5

4.5

5.5

6.5

(b) CP(

(a) Geometry.
0.13

0.35
0.51

0.12

0.49 0.47

0.25

0.4

0.45

0.2

0.13 0.25 0.35

0.1
0.15

0.3

0.25

0.35

0.3

0.4
0.49

0.2

0.2

0.1 0.05

0.09

0.47 0.45

0.49

0.08
0.47
0.45

0.25

0.25

0.08

0.35

0.35
0.25
0.2 0.3
0.15
0.1

0.3

0.2

0.07
0.47

0.4

0.06
0.45

0.45

0.05
0.3

0.25

0.4

0.35

4.5

5.5

6.5

0.04
3

0.25
0.2
0.15
0.1

3.5

4.5

(d) CP(

N=9 free transition


N=0 forced transition at x/c = 0.1

0.02

0.04

0.06

0.08

Cd

(e) Cl

5.5

6.5

10

15

20

clean case.

, )

0.3

0.3

(c) CP(

0.35

0.05
0

0.05

0.35

3.5

0
0.05
0.1
0.15

0.35
0.4

2
1.75
1.5
1.25
1
0.75
0.5
Cl 0.25
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
0

0.1
0.15

0.3

0.11

0.09

0.04
3

= 0.1.

0.12

0.35

0.51

0.1

0.06

for

0.2 0.3

0.11

0.07

0.1

2
1.75
1.5
1.25
1
0.75
0.5
C 0.25
l
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
20

, )

soiled case.

N=9 free transition


N=0 forced transition at x/c = 0.1

15

10

( )

Cd

(f) Cl

Figure 2: Geometry, polar and performance simulation results for the A-26 airfoil.

10 of 16
American Institute of Aeronautics and Astronautics

0.6
N=9 free transition
N=0 forced transition at x/c = 0.1

0.55
0.5
0.45

0.2
0.15
0.1
x/c 0.05
0
0.05
0.1
0.15
0.2
0.1

0.4

t/cmax=0.32

0.3

0.25
0.2
0

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

0.15

1.1

y/c
Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

0.35

0.1
0.05
0
2.5

3.5

4.5

5.5

6.5

(b) CP(

(a) Geometry.
0.25

0.13
0.45

0.12

0.4

0.2

0.35

0.47

= 0.1.

0.12

0.05

0.11

0.05
0.1

0.11
0.3 0.3

0.25

0.1

0.45
0.45

0.4

0.1

0.1
0.15
0.2

0.3

0.09

0.1 0.15
0.2

0.15
0

0.25

0.2

0.25

0.09

0.35

0.08

0.08

0.15

0.07

0.3

0.35

0.06 0.05

0.06

0.1

0.35
0.4

0.4

0.3
0.25

0.04
3

4.5

5.5

(c) CP(

6.5

0.04
3

3.5

4.5

(d) CP(

N=9 free transition


N=0 forced transition at x/c = 0.1

0.02

0.04

0.06

0.08

Cd

(e) Cl

5.5

6.5

10

15

20

clean case.

, )

0.2

0.05

0.35

3.5

0.25

0.15 0.2 0.25

2
1.75
1.5
1.25
1
0.75
0.5
Cl 0.25
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
0

0.1

0.3

0.25

0.4

0.07

0.05

for

0.13

0.1
0.15

0.3

0.1

2
1.75
1.5
1.25
1
0.75
0.5
C 0.25
l
0
0.25
0.5
0.75
1
1.25
1.5
1.75
2
20

, )

soiled case.

N=9 free transition


N=0 forced transition at x/c = 0.1

15

10

( )

Cd

(f) Cl

Figure 3: Geometry, polar and performance simulation results for the A-32 airfoil.

11 of 16
American Institute of Aeronautics and Astronautics

Tangential Force for TSR = 3, pitch = 0

Tangential Force for TSR = 4, pitch = 0


A23 N = 0
A23 N = 9

0.5

0.25
Tangential Force []

Tangential Force []

0.4

0.3

0.2

0.1

0.2

0.15

0.1

0.05

0.1
0

45

90

135

180
225
Azimuth []

270

315

360

(a) A-23 performance for 2 dierent soiling cases at


3.

45

90

135

180
225
Azimuth []

270

315

360

(b) A-23 performance for 2 dierent soiling cases at


4.

Figure 4: A-23 airfoil tangential forces show eect of soiling.


Tangential Force for TSR = 4 and N = 9

Tangential Force for TSR = 3 and N = 9


A26 pitch= 0
A26 pitch= 4

0.6

0.3
Tangential Force []

Tangential Force []

A26 pitch= 0
A26 pitch= 4

0.35

0.5

0.4

0.3

0.2

0.25
0.2
0.15
0.1

0.1

0.05
0

0
0

45

90

135

180
225
Azimuth []

(a) FT for

270

315

360

45

90

= 3.

135

180
225
Azimuth []

(b) FT for

Normal Force for TSR = 3 and N = 9


2.5

270

315

360

= 4.

Normal Force for TSR = 4 and N = 9


2

A26 pitch= 0
A26 pitch= 4

A26 pitch= 0
A26 pitch= 4

1.5

Normal Force []

1.5
Normal Force []

Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

A23 N = 0
A23 N = 9

0.3

0.5

0.5

0
0
0.5
0.5
0

45

90

135

180
225
Azimuth []

(c) FN for

270

315

360

= 3.

45

90

135

180
225
Azimuth []

(d) FN for

Figure 5: A-26 airfoil influence of pitch angle on forces.

12 of 16
American Institute of Aeronautics and Astronautics

= 4.

270

315

360

Tangential Force for TSR = 4 and N = 9


Tangential Force for TSR = 4, pitch = 4
NACA0018 pitch= 0
A26 pitch= 4
0.25

0.25

0.2

0.15

A26 N = 0
A26 N = 4, T=10
A26 N = 4, T=20
A26 N = 9

Tangential Force []

0.15

0.1

0.05

0.1
0.05
0
0.05
0.1
0.15

0
0

45

90

135

180
225
Azimuth []

270

315

0.2

360

(a) Clean performance comparison between A-26 and


NACA 0018.

45

90

135

180
225
Azimuth []

270

315

360

(b) A-26 performance for 3 dierent soiling cases.

Figure 6: A-26 airfoil tangential forces.

Tangential Force for TSR = 4 and N = 9

Tangential Force for TSR = 3 and N = 9


0.6

A32 pitch= 0
A32 pitch= 4

0.5

A32 pitch= 0
A32 pitch= 4

0.3
0.25

0.4
0.3

Tangential Force []

Tangential Force []

Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

Tangential Force []

0.2

0.2
0.1
0

0.2
0.15
0.1

0.1

0.05

0.2

0.3
0.05
0

45

90

135

180
225
Azimuth []

(a) FT for

270

315

360

= 3.

45

90

135

180
225
Azimuth []

(b) FT for

Figure 7: A-32 airfoil influence of pitch angle on forces.

13 of 16
American Institute of Aeronautics and Astronautics

= 4.

270

315

360

Tangential Force for TSR = 4 and N = 9


Tangential Force for TSR = 4, pitch = 0
NACA0018 pitch= 0
A32 pitch= 0

0.3

0.3

0.2

0.2

Tangential Force []

Tangential Force []

0.15

0.1

0.05

0.1

0.1

0.2

0.3
0

45

90

135

180
225
Azimuth []

270

315

360
0

(a) Clean performance comparison between A-32 and


NACA 0018.

45

90

135

180
225
Azimuth []

270

315

360

(b) A-32 performance for 3 dierent soiling cases.

Figure 8: A-32 airfoil tangential forces.

0.6
A23
A26
A32
NACA 0018

0.5

0.4

Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

0.25

A32 N = 0
A32 N = 4, T=10
A32 N = 4, T=20
A32 N = 9

0.3

0.2

0.1

Figure 9: CP

2.5

3
TSR

3.5

curves from the airfoils for the double wake panel model in clean condition (N=9).

14 of 16
American Institute of Aeronautics and Astronautics

1
A23
A26
A32
NACA 0018

0.9

0.8
0.7

0.6

0.6
xtr [x/c]

0.7

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

0
20

15

10

0
[]

10

15

0
20

20

(a) Location of transition point versus angle of attack for


the upper side of the airfoil.

15

10

0
[]

10

15

20

(b) Location of transition point versus angle of attack for


the lower side of the airfoil.

Figure 10: Movement of transition point at Re = 1 million.


Normal Force for TSR = 3 and N = 0

Tangential Force for TSR = 3 and N = 0

Double wake panel model


CFD simulation

2.5

0.3
0.25
Tangential Force []

1.5
Normal Force []

Double wake panel model


CFD simulation

0.35

1
0.5
0

0.2
0.15
0.1
0.05
0

0.5

0.05
1
0.1
0

45

90

135

180
225
Azimuth []

(a) FN for

270

315

360

45

90

= 3.

135

180
225
Azimuth []

(b) FT for

Normal Force for TSR = 4 and N = 0


2

270

315

360

= 3.

Tangential Force for TSR = 4 and N = 0


0.25

Double wake panel model


CFD simulation

Double wake panel model


CFD simulation

0.2

1.5

0.15

Tangential Force []

Normal Force []

Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

xtr [x/c]

0.8

A23
A26
A32
NACA 0018

0.9

0.5

0.1

0.05

0
0.5
0.05
1

0.1
0

45

90

135

180
225
Azimuth []

(c) FN for

270

315

360

= 4.

45

90

135

180
225
Azimuth []

(d) FT for

270

= 4.

Figure 11: A-26 airfoil comparison between CFD and double wake results.

15 of 16
American Institute of Aeronautics and Astronautics

315

360

Normal Force for TSR = 3 and N = 0


Double wake panel model
CFD simulation

Double wake panel model


CFD simulation

0.4

2.5

0.3

0.2
Tangential Force []

Normal Force []

Tangential Force for TSR = 3 and N = 0

1.5
1
0.5
0

0.1
0
0.1
0.2

0.5
0.3
1

45

90

135

180
225
Azimuth []

(a) FN for

270

315

360

45

135

180
225
Azimuth []

(b) FT for

270

315

360

= 3.

Tangential Force for TSR = 4 and N = 0


0.3

Double wake panel model


CFD simulation

90

= 3.

Normal Force for TSR = 4 and N = 0

Double wake panel model


CFD simulation

0.2
1.5
Tangential Force []

0.1
Normal Force []

Downloaded by TECHNISCHE UNIVERSITEIT DELFT on February 20, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.2015-0722

0.4
1.5

0.5

0
0.1
0.2

0.5
0.3
1
0.4
0

45

90

135

180
225
Azimuth []

(c) FN for

270

315

360

= 4.

45

90

135

180
225
Azimuth []

(d) FT for

270

= 4.

Figure 12: A-32 airfoil comparison between CFD and double wake results.

Figure 13: The vorticity field of the A-23 airfoil at

16 of 16
American Institute of Aeronautics and Astronautics

= 4.

315

360

Вам также может понравиться