19
Mini-project report Computational Fluid Dynamics Analysis of a Low Cost Wind Turbine Jon Leary – [email protected] June 2010

Mini Project -MECH

Embed Size (px)

Citation preview

Page 1: Mini Project -MECH

Mini-project report

Computational Fluid Dynamics

Analysis of a Low Cost Wind Turbine

Jon Leary – [email protected]

June 2010

Page 2: Mini Project -MECH

Abstract

Computational Fluid Dynamics (CFD) software was used to compare the performance of a hand-

made wind turbine blade with that of a conventional factory made model. The geometry was

simplified to 2D aerofoils and the surrounding flow field was analysed at a Reynold’s number of

80,000. It was found that the lift/drag characteristics of the two aerofoils across a range of angles of

attack were virtually identical, meaning that the torque force exerted on the wind turbine blades

would also be identical and therefore as would the power outputs of the two turbines. However, the

simple model ignored a number of important issues, such as 3D effects and the influence of

manufacturing quality on the idealised blade geometry. Further modelling and/or experimental

validation work is needed to increase confidence in the quality of the results.

Table of Contents

Abstract ............................................................................................................................................ 2

1 Introduction to Small Scale Wind Power .................................................................................... 3

2 Aims and Objectives .................................................................................................................. 3

3 Basic Aerodynamic Theory of Horizontal Axis Wind Turbines (HAWTs) ....................................... 4

3.1 The Aerofoil ....................................................................................................................... 4

3.2 Lift and Drag, Thrust and Torque ........................................................................................ 5

3.3 Reynold’s Number .............................................................................................................. 8

3.4 Boundary Layers................................................................................................................. 9

3.5 Stall .................................................................................................................................... 9

3.6 Aerofoil Geometry ........................................................................................................... 10

4 Introduction to Computational Fluid Dynamics (CFD) ............................................................... 11

4.1 The Modelling Process ..................................................................................................... 11

4.2 CFD for Wind Turbine Analysis ......................................................................................... 12

5 Building the Model .................................................................................................................. 12

5.1 Modelling Software .......................................................................................................... 12

5.2 Modelling Strategy ........................................................................................................... 12

5.3 Aerofoil Geometry ........................................................................................................... 13

5.3.1 2D Modelling Domain ............................................................................................... 13

5.3.2 Turbulence Model .................................................................................................... 14

6 Results ..................................................................................................................................... 16

7 Analysis ................................................................................................................................... 18

8 Evaluation................................................................................................................................ 18

9 Conclusion ............................................................................................................................... 18

10 References ........................................................................................................................... 19

Page 3: Mini Project -MECH

1 Introduction to Small Scale Wind Power

Small scale wind turbines can be used to provide power

to remote areas of the developing world that are far

away from any existing electrical grid system. The

electricity they supply can be used to provide light in the

mornings and evenings which can allow children to

study and further their opportunities in later life or

adults to continue working and provide that little bit of

extra income for their families that could allow them to

work their way out of poverty. Unfortunately, at a cost

of thousands of pounds, factory built small scale wind

turbines are expensive, even for reasonably well off

citizens of the developed world. Families in the

developing world living below the poverty line often

have to survive on less than $US 1 (~70p) [3] per day,

keeping these potentially revolutionary machines way

out of their reach. However, it is possible to build small

scale wind turbines by hand, using basic workshop tools

and techniques. Designs for such machines are freely

available as open source documents on the internet to

give people in the developing world access to the

technology. Machines such as Hugh Piggott’s [2]

Horizontal Axis Wind Turbine (HAWT) shown in Figure 1 have been tried and tested throughout the

developed and developing world and have proved an invaluable aid to the remote communities in

which they have been installed. At around £500, they cost a fraction of their factory made

counterparts and consequently bring wind power within the reach of isolated small community

groups in the developing world.

2 Aims and Objectives

Piggott’s design may be tried and tested, but has it been truly optimised? Obviously a compromise

has been struck between performance, cost, durability and manufacturability, but is there a way to

improve the performance of the machine without significantly affecting the other factors? This

project aims to:

• Compare the performance of Piggott’s wooden hand carved wind turbine blades with those

of a conventional factory made turbine.

• Propose and evaluate appropriate techniques for improving the efficiency of Piggott’s

turbine blades.

Figure 1 - 1kW wind turbine installed in

Tanzania [2]

Page 4: Mini Project -MECH

3 Basic Aerodynamic Theory of Horizontal Axis Wind

Turbines (HAWTs)

Lift-based Horizontal Axis Wind

Turbines (HAWTs) have today

become the standard

mechanism for harnessing the

power in the wind and

converting it into electricity.

They evolved from the grain

grinding drag-based windmills

of yesteryear. As aeronautics

took off during the last century

with the commercialisation of

aeroplane technology, more

and more became known about

wing technology. The three-

bladed wind turbines we see all

around us today evolved from

aircraft wing theory and are

essentially three wings bolted

onto a generator with a

common central axis. They use

lift from the oncoming wind to

rotate themselves about this

central axis, which points

horizontally into the wind,

hence the name HAWT. Vertical

Axis Wind Turbines (VAWTs)

are also used, but the higher

efficiencies of HAWTs have made them the dominant technology in today’s society.

3.1 The Aerofoil

An aerofoil (or airfoil in the USA) is a 2D shape capable of producing a reactive lift force when in

motion relative to the surrounding air. Most commonly known as the cross-sectional profile of an

aircraft’s wing, Figure 2 shows how the 3D geometry of a lift-based HAWT blade also consists of a

series of aerofoil profiles. Figure 3 gives an overview of the basic aerofoil terminology.

LIF

T

DRAG

Figure 2 - 2D Aerofoil section through a 3D lift-based HAWT blade

(Adapted from [1])

Page 5: Mini Project -MECH

3.2 Lift and Drag, Thrust and Torque

The forces of lift and drag can be easily visualised by considering the force propelling two identical

sailing boats with different types of sails. A boat equipped with a Spinnaker sail, as shown in Figure 4

will be pushed along in the direction of the wind by drag forces. However, a Bermuda rig sail will be

pushed in a direction perpendicular to the wind due to the action of lift forces.

Figure 5 shows how the lift and drag forces are defined on an aerofoil as forces perpendicular and

parallel to the airflow. As the air travels over the top of the aerofoil, it accelerates and consequently

pressure decreases in this area. Lower pressure on the top of the aerofoil than the bottom creates a

suction force called lift. Lift is the force that keeps aeroplanes in the sky. Drag on the other hand,

acts in the same direction to the airflow and is generally considered a nuisance, as for example in an

aeroplane, extra fuel must be used to overcome the drag forces. Drag forces arise mainly from

AIR

FLO

W

LEADING EDGE AVERAGE

CAMBER

LINE

CHORD

TRAILING

EDGE

THICKNESS

ANGLE

OF ATTACK (AoA)

Figure 3 – Aerofoil terminology

Figure 4 – Lift and drag forces propelling a sailing boat

Page 6: Mini Project -MECH

friction between the viscous fluid and the surface of the aerofoil (skin friction drag) and the

difference in pressure between the leading and trailing edges of the aerofoil (form drag).

When comparing different aerofoils, it is often more useful to look at lift and drag coefficients rather

than the total lift and drag forces on acting on the aerofoil. Lift and drag coefficients are non-

dimensional numbers used to quantify the amount of lift or drag on a given aerofoil under a given

set of flow conditions, e.g. Reynold’s number or AoA. They can be defined as follows:

�� ���

12� ��

�� ���

12� ��

CL = Lift coefficient (dimensionless)

CD = Drag coefficient (dimensionless)

ρ = Fluid density (kg/m3)

V = Velocity (m/s)

A = Area of blade (m2), for 2D aerofoil calculations a blade length of 1 is used and only the chord length (m) is required

It should be noted that in the context of a wind turbine, the airflow seen by the aerofoil is not equal

to the oncoming wind velocity. Figure 6 illustrates how the airflow onto the aerofoil (the relative

wind velocity) is a result of both the oncoming wind velocity and the airflow due to blade rotation.

LIFT

DRAG

RESULTANT

FORCE A

IRF

LOW

Figure 5 – Lift and drag forces on an aerofoil

Page 7: Mini Project -MECH

The Tip Speed Ratio (TSR) is a commonly used parameter in wind turbine design and relates the

speed at which the blade tip is travelling (ωr) to the velocity of the oncoming wind:

�� ���

TSR = Tip speed ratio (dimensionless)

ω = Angular velocity of the turbine (Hz)

r = Turbine radius (m), i.e. blade length

V∞ = Wind velocity (m/s)

When calculating the power output of a wind turbine, it is often more relevant to resolve the

resultant aerodynamic force into torque and thrust rather than lift and drag. Torque is the useful

force that causes the blades to rotate, whilst thrust is the redundant component that merely pushes

against the tower. The power produced by a wind turbine can be calculated very simply as follows:

� � �

P = Power (W)

T = Torque (Nm)

ω = Rotational speed (Hz)

It can be clearly seen from Figure 7 that increasing the ratio of lift to drag will result in maximum

torque.

ONCOMING WIND

VELOCITY

DIRECTION

OF BLADE ROTATION

RELATIVE WIND VELOCITY

AIR

FLO

W

AIR VELOCITY DUE

TO BLADE ROTATION

Figure 6 – Velocity triangle relating oncoming wind velocity and blade rotation to the relative

velocity seen by an aerofoil section of a wind turbine blade

Page 8: Mini Project -MECH

3.3 Reynold’s Number

Reynold’s Number is a ratio of the inertial forces to the viscous forces acting within a fluid. It can be

used to characterise flow types: for example laminar flow occurs at low Reynolds numbers, where

viscous forces are dominant, and is characterised by smooth, constant fluid motion, while turbulent

flow occurs at high Reynolds numbers and is dominated by inertial forces, which tend to produce

random eddies, vortices and other flow instabilities. It can be calculated as follows:

LIFT

DRAG

RESULTANT FORCE

AIR

FLO

W

THRUST

TORQUE

THRUST

TORQUE

THRUST

TORQUE

TORQUE THRUST

Figure 7 – Aerodynamic forces on an aerofoil resolved as torque and thrust forces on a wind

turbine (Adapted from [1])

DIRECTION

OF BLADE ROTATION

DIRECTION OF BLADE

ROTATION

Page 9: Mini Project -MECH

�� ��������� ������

������ ��������

! ��

!

Re = Reynold’s number (dimensionless)

ρ = Fluid density (kg/m3)

V = Fluid velocity (m/s)

L = Characteristic length (m), e.g. aerofoil chord length

µ = Dynamic viscosity (kg/ms)

3.4 Boundary Layers

The boundary layer is a thin layer of fluid near the surface of an aerodynamic body which is affected

by the friction between the surface of the body and the viscous fluid flowing around it. This friction

results in the force of skin friction drag. The properties of the boundary layer depend on the

Reynold’s number of the overall flow and on the local conditions. The properties of the boundary

layer around an aerofoil can be grouped together into three categories:

Laminar flow – at low Reynold’s numbers layers of fluid slide smoothly over each other in a thin

boundary layer resulting in low skin friction drag. The flow over an aerofoil will usually start off

laminar before transitioning into turbulent as it travels away from the leading edge. The position of

the transition point is dependent on Reynold’s number and the localised flow conditions.

Turbulent flow – at higher Reynold’s numbers, or after the transition point, the flow near the surface

becomes unstable and the boundary layer thickens, leading to higher skin friction drag.

Separated flow – a high adverse pressure gradient occurring towards the trailing edge of the aerofoil

can cause the flow to separate. Flow near the surface reverses direction and a large turbulent wake

is created from swirling eddy currents that dramatically increases skin friction drag.

Figure 8 – The development of the boundary layer around an aerofoil [4]

3.5 Stall

The lift force produced by a given aerofoil can be increased by raising the angle of attack (AoA), i.e.

lifting the leading edge. At low AoAs, air flows smoothly over the aerofoil and lift increases linearly.

Drag also increases, but not as rapidly as lift, resulting in an increased torque force and therefore

Page 10: Mini Project -MECH

increased power output for the wind turbine. However, as the AoA increases further, the separation

point (which previously had been nestled at the trailing edge, having little influence on the flow

characteristics) jumps forward. Lift reaches a maximum at the stall angle and drag increases

dramatically as the boundary layer becomes separated over the majority of the aerofoil’s upper

surface. Beyond the stall angle, the large turbulent wake greatly reduces lift and as a result a wind

turbine with stalled blades will become both noisy and inefficient.

3.6 Aerofoil Geometry

Aerofoils are used in a variety of applications, from jumbo jet wings to racing car spoilers to wind

turbine blades. For each application, an optimised aerofoil geometry will have been selected to give

to give maximum performance. Figure 10 depicts a number of conventional aerofoil geometries and

their respective characteristics. In general, aerofoils with higher camber tend to give more lift and

although thicker aerofoils have higher drag, they tend to maintain the same level of drag over a

wider range of AoAs and therefore are more stable. For a wind turbine blade, a low speed aerofoil

with high lift to drag ratio would be most desirable.

ANGLE OF ATTACK

X

X

X

X

Figure 9 – The development of stall with increasing angle of attack and its effect on lift.

LIF

T

Page 11: Mini Project -MECH

4 Introduction to Computational Fluid Dynamics (CFD)

Computational Fluid Dynamics (CFD) is a powerful tool used to model the real life behaviour of

fluids. It allows the optimisation of design parameters without the need for the costly testing of

multiple prototypes. What is more, it is also a powerful graphical tool for visualising flow patterns

that can give insight into flow physics that otherwise would be very difficult and costly to discover

experimentally, if possible at all. Governing equations exist to model fluid behaviour, but it is not

always possible to apply them to many of the complex flow patterns we see in the real world directly

as there would be too many unknown variables. However, CFD involves creating a computational

mesh to divide up real world continuous fluids into more manageable discrete sections. The

governing equations for fluid flow can then be applied to each section individually, but as the

properties of each section are inevitably linked to its neighbouring sections, all the sections can be

solved simultaneously until a full solution for the entire flow field can be found. This method

obviously requires a huge amount of computational power, nevertheless with the advancement of

modern computing, solutions that would take months to compute by hand can now be found in

seconds using nothing more than an ordinary desktop or laptop computer.

As with everything, CFD is not without its limitations. Its accuracy or validity are dependent on a

multitude of different factors: the quality and appropriateness of the mesh, the degree to which the

chosen equations match the type of flow to be modelled, the interpretation of the results, the

accuracy of the boundary conditions entered by the user or the level of convergence of the solution,

to name but a few. Often it comes down to the skill of the user, as each flow problem will be slightly

different and as a result, will require a slightly different modelling approach. However, experimental

data can provide a valuable reference point with which to check the validity of CFD models.

4.1 The Modelling Process

The modelling process consists of first taking the real world fluid geometry and replicating this in the

virtual environment. From here, a mesh can be created to divide the fluid up into discrete sections.

Boundary conditions must then be entered into the model to designate parameters such as the type

Figure 10 – Conventional aerofoil geometries, with their characteristics and applications.

Page 12: Mini Project -MECH

of fluids to be modelled or the details of any solid edges or flow inlets/outlets. The simulation is then

ready to be run and when a converged solution is found, it must be carefully analysed to establish

whether the mesh is appropriately modelling the flow conditions. Generally, some form of mesh

refinement will be necessary to put in further detail around the areas of interest.

Figure 11 – The modelling process

4.2 CFD for Wind Turbine Analysis

CFD allows virtual experimentation with and consequently optimisation of the design parameters

such as airfoil shape or angle of attack across a wide range of operating conditions. It is very

attractive to industry as it saves both time and effort during the design process when compared

alongside traditional experimental methods. However, the degree of confidence in the results is

dependent on many factors and as a result; data should be compared with and validated against

experimental findings wherever possible.

5 Building the Model

5.1 Modelling Software

The mesh generation programme Gambit version 2.2.30 was used in conjunction with the solver

Fluent version 6.3.26 to perform the CFD analysis for this project.

5.2 Modelling Strategy

A 2D model of an aerofoil section from the blade tip (Figure 12) was created, with the aim of

calculating lift and drag data for the aerofoil at varying AoAs. This data can give an estimation of the

relative performance of the modelled wind turbines. The aerofoil at the blade tip was chosen as this

is the part of the blade that generates the most lift and therefore its performance is most critical to

the overall power output of the turbine. Varying the AoA simulates the varying wind conditions that

the turbine is likely to experience in service.

Page 13: Mini Project -MECH

Figure 12 – Cross-section of a wind turbine blade showing the 2D aerofoil section to be modelled (turquoise)

and the airflow around it (red)

5.3 Aerofoil Geometry

The NACA0012 aerofoil was modelled first, as the low Reynold’s number experimental data

necessary to validate the model is widely available for this aerofoil in the literature [5]. Once a

satisfactory 2D model had been constructed, the aerofoil of a conventional HAWT blade (LS0413 [6])

and the aerofoil profile from the tip of the Piggott turbine blade would then be modelled and their

performances compared. Figure 13 shows the geometries of the three aerofoils to be modelled.

With reference to Figure 10, the NACA0012 aerofoil is a thin symmetric aerofoil, so zero lift at zero

AoA would be expected. Both the LS01413 and Piggott aerofoils are cambered and as a result, some

lift would be predicted even at zero AoA. The Piggott aerofoil resembles the high speed, low drag

aerofoil at the top of Figure 10; however this does not correlate well with the low speed application

of a wind turbine. As a result, the performance of the LS0413 aerofoil is expected to be superior to

the Piggott aerofoil in the context of wind turbine blades. Specifically, whilst the Piggott aerofoil may

be expected to have lower drag, the LS0413 aerofoil would be expected to create far more lift,

resulting in a much higher thrust force to drive the wind turbine.

Figure 13 – Geometry of the standard NACA0012 aerofoil (top left), conventional wind turbine blade tip

LS0413 aerofoil (top right) and the Piggott turbine blade tip aerofoil (bottom)

5.3.1 2D Modelling Domain

In order to accurately simulate free-stream conditions, a far-field boundary was used at a distance of

at least 12 chord lengths [7] from the aerofoil surface to create the control volume for the analysis.

Page 14: Mini Project -MECH

To model varying AoAs, the free-stream airflow was rotated, whilst the aerofoil remained horizontal.

A wind velocity of 4.5m/s with a TSR of 5.5 was used, giving an inlet velocity of 25.155m/s. As the

geometry was relatively simple, a structured quadrilateral mesh was used to maximise the accuracy

of the model.

5.3.2 Turbulence Model

A number of different turbulence models were suitable candidates for modelling the flow over a 2D

aerofoil. The k-ε RNG and k-ω SST models are both popular choices, however during preliminary

modelling it was found that the S-A (Spalart-Allmaras) model most accurately predicted the lift and

drag characteristics of the NACA0012 aerofoil. The S-A model was designed specifically for low-

Reynold’s number aerospace calculations and has been shown to give good results when simulating

boundary layers subjected to adverse pressure gradients [8]. Figure 15 shows how the k-ε RNG and

S-A models both gave fairly similar predictions of for the lift coefficient up until the stall angle. Due

FAR-FIELD

BOUNDARY

AEROFOIL

AIR

FLO

W

Figure 14 – The computational mesh generated for the CFD analysis

Page 15: Mini Project -MECH

to the turbulent nature of the physical flow conditions that are being modelled close to and beyond

the stall angle, complex time-varying simulations would be required to correctly simulate this

behaviour. As a result, the data obtained from this simple steady-state model in this region cannot

be considered reliable. With regards to the drag coefficient, Figure 15 clearly demonstrates that the

S-A model gives a far better match to the experimental data. As a result, it was decided to use the S-

A turbulence model for the main analysis.

Figure 15 - Comparison of k-ε RNG and S-A turbulence models for the NACA0012 aerofoil for validation

against low Reynold’s number experimental data

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 5 10 15 20

Lift

Co

eff

icie

nt

Angle of Attack

NACA 0012 Model Validation Study

Lift Coefficient

Experimental, Re=160,000 k-E RNG, Re=80,000 S-A, Re=80,000

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0 5 10 15 20 25

Dra

g C

oe

ffic

ien

t

Angle of Attack

NACA 0012 Model Validation Study

Drag Coefficient

k-E RNG, Re=80,000 Experimental, Re=160,000 S-A, Re=80,000

Page 16: Mini Project -MECH

As previously mentioned, accurately modelling the boundary layer is critical for successfully

predicting the drag on an aerofoil. It is necessary to have sufficient mesh in the area adjacent to the

aerofoil surface to accurately resolve the rapidly spatially varying fluid properties within this region.

In order to establish the validity of the model within the boundary layer, a mesh independence study

was carried out. This involved doubling the resolution of the mesh in the region closet to the aerofoil

surface and re-running the model to see the impact on the parameters of interest, in this case, lift

and drag. This is an iterative process and the model is said to be mesh independent when the

parameters of interest no longer change between iterations. Throughout the process, y+ values

around the aerofoil were monitored. y+ is a non-dimensional measure of the height of the cell

adjacent to the aerofoil surface against the height of the boundary layer (and other turbulence

parameters) and is an indicator of how accurately the boundary layer is being modelled. After

successfully conducting the mesh independence study, the y+ values were found to be 1.4 or below

(when using the S-A turbulence model, it is recommended that the values of y+ around the aerofoil

are around 1 or below [8]). Figure 16 graphically demonstrates the ability of the refined model to

resolve the boundary layer sufficiently. Each arrow on the illustration begins in the centre of an

individual cell and it is clear to see that the higher cell density close to the aerofoil surface allows the

correct velocity profile (as seen in Figure 8) to be resolved.

6 Results

Using the validated model, simulations were run of the conventional wind turbine blade tip aerofoil

(LS0413) alongside the Piggott turbine blade tip aerofoil. The angle of attack was varied between 0°

and 15° at a Reynold’s number of 80,000. Figure 17 displays the lift and drag coefficient data for

both aerofoils.

Figure 16 – Vector plot showing the development of the boundary layer around the aerofoil. Colours and

arrow length represent the magnitude of the velocity with arrow direction representing the direction.

Page 17: Mini Project -MECH

Figure 17 – Comparison of the simulated wind turbine blade tip aerofoils

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

0 2 4 6 8 10 12 14 16

Lift

Co

eff

icie

nt

Angle fo Attack (degrees)

Lift Coefficient Comparison

Conventional Wind Turbine Blade Tip (LS01413) Piggott Turbine Blade Tip

0.00

0.02

0.04

0.06

0.08

0.10

0.12

0.14

0.16

0.18

0 2 4 6 8 10 12 14 16

Dra

g C

oe

ffic

ien

t

Angle fo Attack (degrees)

Drag Coefficient Comparison

Conventional Wind Turbine Blade Tip (LS0413) Piggott Turbine Blade Tip

Page 18: Mini Project -MECH

7 Analysis

Figure 17 clearly indicates that the performance of the two wind turbine blade tip aerofoils is

virtually identical. This result is highly unexpected as the geometries of the two aerofoils are very

different. This would seem to suggest that wind turbines using either of these profiles would

produce similar amounts of power. The Piggott aerofoil is a far simpler shape to manufacture, as the

lower surface is effectively a flat surface and therefore it is a far more appropriate design for low-

cost hand manufacturing as virtually no performance is sacrificed.

8 Evaluation

Although the Piggott aerofoil may seem to have an identical performance to the conventional

LS0413 aerofoil, the model used was very simple and a number of factors that will influence the

performance of a wind turbine blade in real life were not included, for example:

• 3D effects – Real wind turbine blades are both twisted and tapered and encounter

increasing relative wind speeds towards the tip due to the rotation of the blade. Vortices are

also created at the ends of the blades due to the difference in pressure between the upper

and lower surfaces of the aerofoil. In order to correctly model the geometry of a wind

turbine blade, a 3D model would need to be built.

• Transition point – The point along the surface of the aerofoil at which the flow transitions

from laminar to turbulent is of critical importance to determining the drag on the aerofoil.

To date, no computational techniques are capable of predicting the location of this point

and experimental measurements must be taken to determine its location. As a result, the

entire boundary layer was modelled as turbulent and the drag is likely to be an over-

prediction.

• Idealised geometry – The wind turbine blade tips were modelled as ideal aerofoil

geometries, however this would not be the case in the real world. In particular, a hand-made

wind turbine blade is likely to be full of defects arising from poor manufacturing technique,

especially at the tip where the size of the aerofoil relative to the size of the tools is smallest.

These defects, although they may seem small, could have a critical impact on the aerofoil’s

performance as they could trip the flow from laminar to turbulent and consequently

increase drag. The surface roughness of the wind turbine blades was also not modelled,

which could also similarly affect the drag.

• Accuracy of CFD – As shown by Figure 15, the model still has a significant degree of

inaccuracy, especially near and above the stall angle. Further modelling is required to

increase the accuracy of the model, in particular unsteady simulations to more accurately

determine performance around stall angle.

• Experimental validation – Although the model was validated using the NACA0012 aerofoil,

the flow physics may be slightly different around the two test aerofoils and as a result

experimental data for these would give more confidence in the results.

9 Conclusion

Hugh Piggott’s DIY wind turbine has been shown to have comparable performance to that of a

conventional wind turbine. In the simplified model of the aerofoils at the blade tips, both exhibited

virtually identical lift and drag characteristics, implying that the torque force exerted on the blades

Page 19: Mini Project -MECH

and consequently the power produced by the turbine would also be identical. However, the simple

model neglects many important factors such as 3D effects, geometric variations deriving from

manufacturing defects and the location of the transition point. As a result, further modelling and/or

experimental work is required to give more confidence in the results of this study.

10 References

[1] How Stuff Works.com. Available: www.howstuffworks.com, Accessed 18th July 2010

[2] H. Piggott. (2010, Scoraig Wind. Available: www.scoraigwind.com

[3] A. Doig, "Off-grid Electricity for Developing Countries," IEEE Review, pp. 25-28, 1999.

[4] 16th July). My Aviation Tutor. Available: http://myaviationtutor.com

[5] R. E. a. K. Sheldahl, P. C., "Aerodynamic Characteristics of Seven Airfoil Sections Through 180

Degrees Angle of Attack for Use in Aerodynamic Analysis of Vertical Axis Wind Turbines,"

Sandia National Laborotories, Albuquerque, New Mexico, USA1981.

[6] K. Kishinami, et al., "Theoretical and Experimental Study on the Aerodynamic Characteristics

of a Horizontal Axis Wind Turbine," Elsevier, 2005.

[7] R. Bhaskaran. 1st July). Cornell Fluent Tutorials - Flow Over an Airfoil. Available:

http://courses.cit.cornell.edu/fluent/airfoil/index.htm

[8] Fluent_Inc. (2005). Fluent 6.3.26 User Manual.