Fourie AZ 2016

Download as pdf or txt
Download as pdf or txt
You are on page 1of 189

Aerodynamic analysis and design of a

solar powered vehicle using CFD

A.Z. Fourie
21588635

Dissertation submitted in fulfilment of the requirements


for the degree Magister in Mechanical Engineering at the
Potchefstroom Campus of the North-West University

Supervisor: Dr J.J. Bosman

May 2016
Abstract

The need for solar-powered vehicles is increasing as the threat of depletion of non-renewable energy

sources, such as oil, increases. This gives reason for developing electric vehicles powered by renewable

energy sources, such as solar power. An important aspect for the overall efficiency of such vehicles is the

aerodynamic design. To improve the design, proper analysis is essential. A generically based methodology

describes the design process applied to improve the aerodynamic performance of these vehicles using

Computational Fluid Dynamics (CFD) as the design tool. CFD is a valuable tool as it accelerates the design

process within a 3-D flow environment. The literature survey conducted during this study describes basic

fluid properties relevant to the flow around a body. It also describes the different turbulence models used

in CFD and the appropriate transition model to be used. Finally, the factors of the body of a solar-powered

vehicle that can be optimised to improve the aerodynamics are also discussed. The methodology followed

is an iterative process comprising of the improvement of the baseline design. It was found that this iterative

method enables the designer to improve the design by analysing various vehicle revisions using CFD and

adjusting certain aerodynamic factors described in the literature. The aim is to minimise the design

parameters with each iteration; in other words the frontal area of the vehicle as well as the drag coefficient.

Other factors greatly affecting the drag coefficient are the junction fillets of the different components as

well as the positions of these components relative to each other. By varying these factors, the near optimum

values could be correlated with the theoretical values and based on this the design could be improved.

Proper airfoil designs should also be implemented when designing the components to improve the laminar

flow. The results found illustrate that the process followed for the analysis and design of the vehicle provide

a suitable method for related problems in future designs. CFD is also determined to be a suitable design

tool. The end result can, however, be validated experimentally, as an assumption was made that the CFD

simulations were adequate, as shown by the results of the validation process. Comparing the simulation

results with experimental results will further prove the validity of the study.

i
Acknowledgements

Firstly all praise to my heavenly Father for the strength He has given me during the duration of this study.

I would like to thank my family – parents Merwe and Christa, brothers Liaan and Stephan and sister-in-law

Michelle – for their continued support and motivation throughout the duration of this study, and all my

friends who also supported and motivated me.

Thanks are also due to:

Dr Johan Bosman for his valuable advice as supervisor and for the financial support required to complete

this study.

Prof Jat Du Toit for help with formulating the title.

Christiaan de Wet (Aerotherm) for hours spent aiding me with countless simulations and valuable advice

in this regard.

Thabo Molambo for the support and use of the HPC (NWU High Performance Computer cluster).

Kitso Epema (TU Delft’s Human Power Team) for the information on the HPT’s VeloX1 geometry and

wind tunnel test results.

Finally, to Arno de Beer and the team for the time spent building the actual vehicle and competing in the

Sasol Solar Challenge.

ii
Table of Contents

Abstract ........................................................................................................................................................ i

Acknowledgements ..................................................................................................................................... ii

Table of Contents ...................................................................................................................................... iii

List of figures ............................................................................................................................................ vii

List of tables ............................................................................................................................................... xi

Nomenclature ............................................................................................................................................ xii

Chapter 1: Introduction............................................................................................................................ 1

1.1 Background .................................................................................................................................. 1

1.2 Problem Statement ....................................................................................................................... 2

1.3 Research Aims and Objectives ..................................................................................................... 3

1.4 Significance .................................................................................................................................. 3

1.5 Definition of key terms ................................................................................................................. 4

Chapter 2: Literature Survey ................................................................................................................... 5

2.1 Aerodynamic Drag ....................................................................................................................... 6

2.1.1 Drag Components ................................................................................................................. 7

2.2 The Boundary Layer ................................................................................................................... 11

2.2.1 Laminar and Turbulent Flow .............................................................................................. 12

2.2.2 Separation ........................................................................................................................... 14

2.3 Transition Modelling with CFD ................................................................................................. 15

2.3.1 The γ-Reθ Transition Model ............................................................................................... 16

2.4 CFD as a Design Tool ................................................................................................................ 18


iii
2.5 Aerodynamic Design .................................................................................................................. 20

2.5.1 Vehicle Body Shape ........................................................................................................... 20

2.5.2 Nose design ........................................................................................................................ 24

2.5.3 Interference Drag Reduction .............................................................................................. 25

2.5.4 Canopy Drag Reduction ..................................................................................................... 29

2.5.5 Trailing Edge Thickness of the Main Body........................................................................ 31

2.6 Popular Solar-powered Vehicles................................................................................................ 32

2.7 Summary ..................................................................................................................................... 37

Chapter 3: Methodology ......................................................................................................................... 38

3.1 Review Regulations and Evaluate Existing Solutions ................................................................ 39

3.1.1 Summary of previous work ................................................................................................ 39

3.2 Geometric Specifications and Initial Design Parameters .......................................................... 39

3.3 Create Concept Designs ............................................................................................................. 41

3.4 Selection of Best Concept Design ............................................................................................... 42

3.5 Detailed Design .......................................................................................................................... 42

Chapter 4: 2-D Validation: NACA 0018 .............................................................................................. 44

4.1 Validation Setup ......................................................................................................................... 44

4.1.1 Mesh Configuration ............................................................................................................ 45

4.1.2 Solver Configuration .......................................................................................................... 48

4.2 Validation Procedure ................................................................................................................. 49

4.3 Validation Results....................................................................................................................... 50

4.4 Conclusion .................................................................................................................................. 57

Chapter 5: 3-D Validation: HPT VeloX1 ............................................................................................. 58

5.1 Validation Setup ......................................................................................................................... 59


iv
5.1.1 Mesh Configuration ............................................................................................................ 60

5.1.2 3-D Solver Configuration ................................................................................................... 62

5.2 Validation Procedure ................................................................................................................. 62

5.3 Validation Results....................................................................................................................... 63

5.4 Conclusion .................................................................................................................................. 68

Chapter 6: Conceptual Design ............................................................................................................... 69

6.1 Concepts ..................................................................................................................................... 69

6.1.1 Concept based on Symmetry .............................................................................................. 70

6.1.2 Asymmetrical Concept with Long Fairings ....................................................................... 70

6.1.3 Asymmetric Design with Short Fairings ............................................................................ 72

6.2 Comparison Setup ...................................................................................................................... 74

6.3 Comparison Procedure .............................................................................................................. 74

6.4 Comparison Results.................................................................................................................... 75

6.5 Conclusion .................................................................................................................................. 78

Chapter 7: Detail Design......................................................................................................................... 79

7.1 Component Analysis ................................................................................................................... 79

7.1.1 Analysis Procedure ............................................................................................................. 79

7.1.2 Main Vehicle Body ............................................................................................................ 79

7.1.3 Fairing Profile Selection ..................................................................................................... 82

7.1.4 Fairing Design .................................................................................................................... 86

7.1.5 Canopy Design ................................................................................................................... 92

7.2 CFD Analysis of Detail Design .................................................................................................. 99

7.2.1. Model Based on Near Optimum Values ........................................................................... 102

7.2.2. Effect of Increasing Fillet Radii on Fairings .................................................................... 105

7.2.3. Effect of Reducing Fillet Radii to Below Theoretical Optimum ...................................... 107
v
7.2.4. Applying the Fillet Radius Theory to Canopy Junctions ................................................. 109

7.2.5. Effect of a Large Canopy Junction Fillet.......................................................................... 110

7.2.6. Effect of Diverting Flow around the Rear Fairing ........................................................... 110

7.2.7. Applying Fillet Radius Theory to the Driver Cockpit Fairing ......................................... 113

7.2.8. Effect of Removing Sharp Floor Edges on Fairings......................................................... 114

7.3 Detail Design Results ............................................................................................................... 116

7.4 Conclusion ................................................................................................................................ 117

Chapter 8: Conclusion and Recommendations .................................................................................. 119

8.1. 2-D Validation .......................................................................................................................... 119

8.2. 3-D Validation .......................................................................................................................... 119

8.3. Detail Design............................................................................................................................ 120

8.4. Recommendations for Future Work.......................................................................................... 120

Bibliography ........................................................................................................................................... 122

Annexure A: 2-D Validation Summary Report .............................................................................. 128

Annexure B: 3-D Validation Summary Report .............................................................................. 138

Annexure C: Detail Design Analysis Summary Report ................................................................. 148

Annexure D: Final Remarks............................................................................................................. 169

vi
List of figures

Figure 1 - Solar-powered vehicles (Tour de Sol 1987) [4] ....................................................................... 5

Figure 2 - Aerodynamic drag and rolling resistance against vehicle speed [7] ..................................... 7

Figure 3 - Comparison between (a) a bluff body and (b) a streamlined body [8]................................. 8

Figure 4 - Wake produced by different body shapes [10] ....................................................................... 9

Figure 5 - Induced drag and lift [12] ...................................................................................................... 10

Figure 6 - Boundary layer along a thin flat plate [14]........................................................................... 12

Figure 7 - Natural transition process [15] .............................................................................................. 14

Figure 8 – Schematic of flow separation at a wall [17] ......................................................................... 15

Figure 9 - Summary of how camber effects the lift force on the body in freestream and in close

ground proximity [5] ........................................................................................................................ 21

Figure 10 - Tapering of the trail edge of a bluff body (boat-tailing) [30] ............................................ 22

Figure 11 - Body at trim [32] ................................................................................................................... 23

Figure 12 - Airfoil contribution to longitudinal stability [33] .............................................................. 23

Figure 13 - Junction flows between the fairing and the body [35] ....................................................... 26

Figure 14 - Inclination angle of an appendage [5] ................................................................................. 27

Figure 15 - Fillet radius along an appendage [5] ................................................................................... 28

Figure 16 - The upper and lower bounds of the canopy drag as a function of the L/h ratio [5] ....... 30

Figure 17 - (a) Illustration of baseline canopy with L/h = 6. (b) Base line canopy with a blunt nose.

(c) Baseline canopy with a blunt tail. (d) Baseline canopy with a blunt nose and tail [5] .......... 31

Figure 18 - Trailing edge thickness plotted against the percentage increase in drag area [5] .......... 32

Figure 19 - Nuna6 solar vehicle [43] ....................................................................................................... 33

Figure 20 - Nuna7 solar vehicle [44] ....................................................................................................... 34

Figure 21 - Tokai University solar-powered vehicle [45] ...................................................................... 35

Figure 22 - Solar Team Twente's "The Red Engine" solar-powered vehicle [42].............................. 36

Figure 23 - Domain with boundaries defined......................................................................................... 45

Figure 24 – Quadrilateral 2-D mesh ....................................................................................................... 46

vii
Figure 25 - Overset mesh: airfoil rotated to an angle of attack of 8° ................................................... 47

Figure 26 - Prism layer on the leading edge of the NACA 0018 airfoil ............................................... 47

Figure 27 - Mesh refinement for the 2-D simulation at an angle of attack of 0° ................................ 49

Figure 28 - Lift and Drag coefficients, 8° angle of attack ..................................................................... 51

Figure 29 - Lift coefficient compared with the Drag coefficient, Re = 0.7x106 ................................... 52

Figure 30 - Lift coefficient compared with the angle of attack, Re = 0.7x106 ..................................... 53

Figure 31 - Separation point on airfoil ................................................................................................... 54

Figure 32 - Skin friction coefficient plot for a 2-D NACA 0018 airfoil, 0° angle of attack ................ 55

Figure 33 - Pressure coefficient of the 2-D NACA 0018 airfoil, 8° angle of attack ............................. 56

Figure 34 - Separation bubble formation on the upper airfoil surface, 8° angle of attack ................ 56

Figure 35 – Human Power Team’s VeloX1 [57] .................................................................................... 58

Figure 36 - Domain description for the 3-D validation case ................................................................. 60

Figure 37 - VeloX1 geometry ................................................................................................................... 61

Figure 38 - Mesh refinement for VeloX1 trailing edge flow ................................................................. 61

Figure 39 - VeloX1 prism layer on surface mesh................................................................................... 62

Figure 40 - STAR-CCM+ force monitor plot ........................................................................................ 64

Figure 41 - VeloX1 transition location with the ventilation ducts taped [57] ..................................... 65

Figure 42 – VeloX1 transition location by VSAero [57]........................................................................ 66

Figure 43 - VeloX1 transition location by STAR-CCM+, turbulent kinetic energy plot ................... 66

Figure 44 - HPT VeloX1 pressure coefficient by VSAero ..................................................................... 67

Figure 45 - VeloX1 pressure coefficient by STAR-CCM+.................................................................... 67

Figure 46 – First concept design (symmetrical) ..................................................................................... 70

Figure 47 – Two long fairings with driver to the right hand side of the vehicle ................................. 71

Figure 48 – Front view of the second concept ........................................................................................ 72

Figure 49 – Asymmetric design with three fairings .............................................................................. 73

Figure 50 – Front view of the third design, showing the effect that the shorter fairings has on the

frontal area........................................................................................................................................ 73

Figure 51 - Turbulent kinetic energy of the first concept ..................................................................... 75


viii
Figure 52 - Bottom view of the symmetric concept vehicle................................................................... 76

Figure 53 - Turbulent kinetic energy of the belly section of the second concept ................................ 77

Figure 54 - Turbulent kinetic energy of the belly section of the third concept ................................... 78

Figure 55 – Modified JS1 sailplane wing profile used for main body ................................................. 80

Figure 56 - Pressure coefficient curves of main body profile ............................................................... 80

Figure 57 - Main body of the vehicle ...................................................................................................... 81

Figure 58 - Rounded leading-edge of the main body............................................................................. 82

Figure 59 - Laminar flow on leading-edge of main body ...................................................................... 82

Figure 60 - RUD20 airfoil ........................................................................................................................ 83

Figure 61 - YS900 airfoil .......................................................................................................................... 83

Figure 62 - NACA 16015 airfoil .............................................................................................................. 84

Figure 63 - NACA 16015 and YS900 airfoils combination ................................................................... 85

Figure 64 - Pressure distribution along the chord length of the NACA 16015 YS900 airfoil............ 85

Figure 65 – Fairing Configuration 1 ....................................................................................................... 86

Figure 66 - Transition locations of flow on fairing Configuration 1 .................................................... 87

Figure 67- Pressure coefficient plot of Configuration 1 ........................................................................ 87

Figure 68 – Fairing Configuration 2 ....................................................................................................... 88

Figure 69 - Transition locations of flow on fairing Configuration 2 .................................................... 89

Figure 70 - Pressure coefficient plots of configuration 2 ...................................................................... 90

Figure 71 – Fairing Configuration 3 ....................................................................................................... 91

Figure 72 - Transition locations of flow on fairing Configuration 3 .................................................... 91

Figure 73 - Pressure coefficient plots of configuration 3 ...................................................................... 92

Figure 74 - Canopy 1 turbulent kinetic energy plot .............................................................................. 93

Figure 75 - Pressure coefficient plot of canopy 1 ................................................................................... 94

Figure 76 - Canopy 2 turbulent kinetic energy plot .............................................................................. 95

Figure 77 - Pressure coefficient plot of canopy 2 ................................................................................... 96

Figure 78 - Canopy 3 turbulent kinetic energy plot .............................................................................. 97

Figure 79 - Pressure coefficient plot of Canopy 3 .................................................................................. 98

ix
Figure 80 – Mesh independence ............................................................................................................ 100

Figure 81 - Detail design simulation boundary definitions ................................................................. 101

Figure 82 - Laminar flow on upper part of the baseline model ......................................................... 103

Figure 83 - Laminar flow on right side of the baseline model ............................................................ 103

Figure 84 - Laminar flow on lower part of the first vehicle model .................................................... 104

Figure 85 - Laminar flow on upper surface of the model after changing the fillet radii to 60 mm 106

Figure 86 - Transition at fairing leading-edges of the revised model ................................................ 107

Figure 87 - Transition at fairing leading-edges of the 30 mm fillet radii model ............................... 108

Figure 88 - Canopy wake of the present model ................................................................................... 109

Figure 89 - Laminar flow on the sixth model ....................................................................................... 111

Figure 90 - Laminar flow on inside of fairings .................................................................................... 112

Figure 91 - Fairing wake due to the modified front fairing ................................................................ 112

Figure 92 – Increase in frontal area due to larger fillets on the right hand fairing (left side in above

image) .............................................................................................................................................. 113

Figure 93 - Rounded lower edges of fairings........................................................................................ 114

Figure 94 - Improved laminar flow on lower edge of fairings ............................................................ 115

Figure 95 - Improved laminar flow on rear left fairing ...................................................................... 115

Figure 96 - Drag coefficients of the vehicle model when moving relative to the ground ................. 117

Figure 97 - NWU Solar Car competing in the Sasol Solar Challenge 2014 ...................................... 169

Figure 98 - Sirius X25 headed to Kroonstad on Day 1 of the Sasol Solar Challenge 2014 .............. 171

Figure 99 – Sirius X25 travelling on the N1 highway between Kroonstad and Bloemfontein, Free

State. ................................................................................................................................................ 171

Figure 100 – On route to Hanover in the Karoo.................................................................................. 172

Figure 101 - Sirius X25 completing a loop in the Karoo ..................................................................... 172

Figure 102 - The vehicle on its way to Port Elizabeth, Eastern Cape ................................................ 173

Figure 103 - Doing a loop between Heidelberg and Witsand in the Western Cape ......................... 173

Figure 104 - Control stop in Heidelberg, Western Cape .................................................................... 174

Figure 105 - Sirius X25 and team at the finish in Cape Town............................................................ 174

x
List of tables

Table 1 - Definition of key terms ............................................................................................................... 4

Table 2 - Relative sizes of the drag components and variables to which each is proportional [5] .... 11

Table 3 - Atmospheric Conditions .......................................................................................................... 48

Table 4 - Drag Area Results Comparison .............................................................................................. 65

Table 5 - Conceptual design comparison................................................................................................ 75

Table 6 - Drag force results of the stationary vehicle.......................................................................... 116

Table 7 - Drag force results of the moving vehicle .............................................................................. 116

xi
Nomenclature

A Frontal area [m2]

c Chord length [m]

CDA Aerodynamic drag area [m2]

CD Aerodynamic drag coefficient [-]

CL Lift Coefficient [-]

Cm Pitching Moment Coefficient [-]

Cp Pressure Coefficient [-]

Cd,frontal Drag coefficient based on frontal area [-]

∆Cd Change in drag coefficient [-]

∆Dj Drag increase due to junctions [N]

D Aerodynamic drag force [N]

Dind Induced drag force [N]

Dint Interference drag force [N]

Dp,sep Pressure drag force [N]

Dp,BL Boundary layer drag force [N]

Dskin Skin drag force [N]

δ Boundary layer thickness [m]

h Height [m]

xii
L Length [m]

Ploc Local pressure [Pa]

P∞ Pressure outside the boundary layer [Pa]

ρ Air density [kg/m3]

Re Reynolds number

Reθt momentum thickness Reynolds number

t Maximum thickness of body airfoil [m]

tte Thickness of trailing edge [m]

U∞ Free stream velocity [m/s]

ν Kinematic viscosity [m2/s]

V Flow velocity [m/s]

Vloc Local velocity [m/s]

V∞ Velocity outside the boundary layer [m/s]

x Distance from leading-edge [m]

Acronyms

CAD Computer Aided Design

CFD Computational Fluid Dynamics

CPU Central Processing Unit

DNS Direct Numerical Simulation

xiii
HPT Human Power Team

LCTM Local Correlation-based Transition Model

LES Large Eddy Simulation

NACA National Advisory Committee on Aeronautics (USA)

RANS Reynolds Averaged Navier-Stokes

SST Shear Stress Transition model

SV Solar-powered Vehicle

Keywords

Computational Fluid Dynamics, CFD, aerodynamics, aerodynamic drag, analysis and design, solar-

powered vehicle, drag coefficient, STAR-CCM+

xiv
Chapter 1: Introduction

1.1 Background

Road vehicles are one of the primary transportation systems and are largely powered by oil, a non-

renewable source of energy. The threat of depleting non-renewable energy sources has encouraged research

into finding alternative sources of energy. Converting solar energy into electricity is such an alternative

source. This has induced the application of solar energy to power electric vehicles.

The aerodynamic drag component of a vehicle’s design plays an important role in its performance and has

to be optimised [1]. Solar panels on typical solar-powered race vehicles usually generate about 1500 W of

electricity, equivalent to the amount used by a hair drier. For this energy to be used effectively, the drag

resistance of the vehicle should be optimised to be as low as possible. It therefore encourages research to

be conducted in optimising the aerodynamic designs of these solar-powered vehicles.

A wind tunnel as a design tool is essential in the aerodynamic development of road vehicles, including

solar-powered vehicles, but building models to test in a wind tunnel is time consuming and expensive. A

wind tunnel in this case may not always be an economical and viable tool. Computational Fluid Dynamics

(CFD) can be used as an alternative to wind tunnel testing. CFD codes simulate detailed fluid flow

problems and allow these problems to be analysed during the early stages of the detail design [2]. This can

possibly result in reduced costs and early detection and avoidance of possible problems during the design

and manufacturing processes.

CFD was therefore used to analyse the baseline design for a solar-powered vehicle (SV) designed at the

North-West University. The results of the first generation SV were used as the baseline from which the

next generation SV was designed. The vehicle is designed and built in order to compete in the Sasol Solar

Challenge. Both local and international solar car teams compete in the South African Sasol Solar Challenge

endurance competition held on the South African national roads.

1
New regulations in the competition meant that the aerodynamic design of the first generation vehicle had

to be modified. These rules included a change in the vehicle dimensions as well as requiring a larger canopy

for increased driver comfort.

1.2 Problem Statement

With recent developments in photovoltaic solar power technology, solar cell performance has increased

significantly, reaching efficiency values as high as 40%. However these efficiencies come at a price and

are therefore too expensive for most applications.

For increased endurance performance of the solar car, one of the options currently available is the

improvement of the aerodynamic shape of the vehicle.

The current (first generation) solar car design of the NWU has typical features that increase the aerodynamic

drag significantly. These features include a swing-arm rear suspension that protrudes from the body. This

increases the frontal area and which increases the overall drag coefficient.

Solar vehicles which performed well in the past, have drag coefficient values of less than 0.1. The current

solar vehicle design has a drag coefficient of almost 0.25, meaning that it is much less efficient than the

current world leaders’ designs. The teams that consistently compete for the number one position at the

World Solar Challenge every other year, complete the challenge at an average speed in excess of 80 km/h,

with top speeds of about 110 km/h. During the previous solar challenge, the current solar vehicle could

only maintain an average speed of about 35 km/h and maintain a top speed of 110 km/h for a very limited

time (less than one hour). The current design of the NWU can be improved in many areas, each of which

is a study in its own. The most significant improvement can be achieved by improving the aerodynamic

design.

In order to improve the aerodynamics of the current vehicle, the complete 3-D flow field around the vehicle

has to be analysed. Analysing this 3-D flow field can show how the combination of the different body

components increases the drag of the vehicle. This can be in terms of reduced laminar flow regions on the

body due to increased interference and poor body orientation. Using CFD, the aerodynamic drag

coefficients of the solar vehicle can be determined by calculating the aerodynamic force on the surface of

2
the vehicle. In order to correctly determine these drag coefficients, transition locations should also be

calculated accurately as both laminar and turbulent flow regions are present at these operating Reynolds

numbers. If transition is predicted incorrectly, the drag coefficients will also be inaccurate.

Recently transition models have become readily available in CFD software packages. However, before

these transition models can be used to design new aerodynamic shapes, the accuracy of the numerical

simulations should be determined. In determining the adequacy of CFD as a design tool for solar vehicles,

a validation of the transition models must be conducted.

1.3 Research Aims and Objectives

 Validate CFD transition model for flow environments similar to solar vehicles.

 Optimise the aerodynamic design of a solar-powered vehicle within the regulations as specified by

the Sasol Solar Challenge, specifically section 1 and section 2 [3], using CFD. The main factor to

consider is the aerodynamic drag of the vehicle. A 5% reduction in the drag coefficient is deemed

notable, if a near optimum base model (based on theory) is selected. An improvement in excess of

10% is deemed acceptable when the new design is compared to the existing vehicle design.

1.4 Significance

Many vehicle manufacturers such as Toyota and Nissan have established a market for electric hybrid

vehicles. These vehicles typically have a small internal combustion engine, only used for long distance

travelling or when the power supplied by the electric motor is exceeded. The vehicle will have a battery

storage system to supply energy to the electric motors and will usually be charged using electricity from a

grid. Many manufacturers also use solar arrays to supply electricity for applications within the vehicle such

as the air conditioning and navigation systems.

The research conducted in this dissertation can therefore further the knowledge of building efficient and

affordable road vehicles powered by solar energy. There is also the option to build a hybrid vehicle, but

instead of charging the batteries from grid electricity, the batteries will be charged by the solar arrays.

Further research will allow these production line vehicles to run solely on solar power, without any internal

combustion engines.
3
1.5 Definition of key terms

The key terms used in this project are summarised in Table 1.

Table 1 - Definition of key terms

A division within fluid mechanics using

Computational Fluid Dynamics (CFD) numerical methods and algorithms to solve

and analyse fluid flow problems.

Designs done using a computer to represent

Computer Aided Design (CAD) a model of the part or vehicle as it will be

constructed.

Refers to the solar-powered vehicles

Solar-powered Vehicle (SV) constructed as a result of the research

conducted in this project.

A team of students form TU Delft in the

Netherlands, who built a bicycle with which


Human Power Team (HPT)
they intend to break the land speed record

for a human powered vehicle

4
Chapter 2: Literature Survey

Solar car racing first started with the Tour de Sol competition held in 1985, leading to other similar

competitions worldwide [4]. These competitions were mainly aimed at university students to find

innovative ways to expand their skills and develop solar technology.

Figure 1 - Solar-powered vehicles (Tour de Sol 1987) [4]

There are many different types of solar-powered cars being built by different institutions, such as the one

in Figure 1. This was one of the first solar-powered racing vehicles built in 1987. Contestants from all

sectors are allowed to participate, from privateers building vehicles out of the cheapest materials available

to highly funded institutions such as universities having access to the latest technologies in order to build

the most efficient vehicles possible.

Two methods are generally applied in the reduction process of the aerodynamic drag, namely the ground-

up approach and the improvement approach [5]. The ground-up approach is where the main body is

designed for lower drag coefficients and the other components are then designed within the body

constraints. The improvement approach is where the designer will improve the features of an already non-

aerodynamic vehicle. These processes will ensure a reduction in the drag of the vehicle.

5
2.1 Aerodynamic Drag

An object with a higher air resistance will require more energy to be displaced. Aerodynamic drag is one

of the main factors adversely affecting the efficiency of a road vehicle or a body moving through air, and

should therefore be optimised. The remainder of the resistance will be due to the rolling resistance caused

by the friction between the tyres and the road surface [5].

The aerodynamic drag of a vehicle is determined by the frontal area (A), the drag coefficient (CD), the

square of the sum of the vehicle’s velocity (Vcar) and the headwind velocity (Vhead) and the air density (ρ).

The drag coefficient of the vehicle is also described as a measure of the flow quality around the vehicle.

According to Tamai [5], aerodynamic drag (D) can be expressed as:

1
𝐷 = (𝐶𝐷 𝐴) 𝜌(𝑉𝑐𝑎𝑟 + 𝑉ℎ𝑒𝑎𝑑 )2 2. 1
2

Vcar is the velocity of the car relative to the road and V head is the velocity of the headwind relative to the

road. The sum of these velocities will be the velocity of the air relative to the car (V). The above equation

can thus be simplified to the following equation, as described by Smith [6]:

1
𝐷 = (𝐶𝐷 𝐴) 𝜌(𝑉)2 2. 2
2

In order to reduce the drag of the vehicle, one has to analyse this equation. The density of the air, ρ, has an

effect on the drag of the vehicle but one cannot adjust it to reduce the drag of the body. The velocity, V at

which the vehicle is travelling, is always kept at a constant value which will deliver the best performance,

but the other two factors do have an effect. By decreasing the frontal area, there is a direct effect on the

drag of the vehicle as they are directly proportional to one another. One can further reduce the drag of the

vehicle by designing a more aerodynamic shape that moves easier through air, thus with a lower CD value.

The total drag of the vehicle, called the drag force, will be the sum of the aerodynamic drag and the rolling

resistance. Refer to Figure 2 to see the relationship between the aerodynamic drag and the rolling resistance

at different speeds [7].

6
Figure 2 - Aerodynamic drag and rolling resistance against vehicle speed [7]

The drag forces will increase as the vehicle speed increases, with the aerodynamic drag having the largest

effect.

2.1.1 Drag Components

The aerodynamic drag (D) of the body has several components, namely pressure drag or form/profile drag

(Dp,sep), viscous friction drag or surface drag (Dskin + Dp,BL), induced drag (Dind) and interference drag (Dint).

According to Tamai [5], the aerodynamic drag can thus be expressed as:

𝐷 = 𝐷𝑝,𝑠𝑒𝑝 + (𝐷𝑠𝑘𝑖𝑛 + 𝐷𝑝,𝐵𝐿 ) + 𝐷𝑖𝑛𝑑 + 𝐷𝑖𝑛𝑡 2. 3

where Dskin is the drag caused by the skin friction and Dp,BL is the pressure drag caused by the boundary

layer, the sum of which makes up the viscous friction drag.

Pressure Drag

Pressure drag occurs when the air flow around the body does not stay attached to the body all the way to

the trailing edge of the vehicle. This phenomenon is called flow separation. When this does occur, a low-

7
pressure wake develops behind the vehicle. This causes a sucking force, pulling the vehicle backwards.

This also results in vortices behind the body, which consumes large amounts of energy. A body in which

separation occurs, is called a bluff body. Pressure drag, or separation pressure drag, is induced by surface

irregularities (surface or skin friction) which causes the flow to separate from the body. The frontal area of

the vehicle also has an effect on the separation of flow.

Figure 3 - Comparison between (a) a bluff body and (b) a streamlined body [8]

Referring to Figure 3 [8], it is clear that (a) a bluff/blunt body has much more pressure drag than (b) a

streamlined body due to the larger wake pulling the object backwards.

Viscous Friction Drag

The viscous friction drag, also called friction drag, of the body has two components. The first of these is

the drag caused by the skin friction, which is the viscous shearing of fluid molecules tangential to the

surface.

In aerodynamic applications, flow separation is not ideal and the design should be such that the airflow

remains attached for as long as possible. A body in which there is no, or very little, flow separation present

8
is called a streamlined body. It should be noted that a bluff body has five to ten times the aerodynamic drag

of a streamlined body [9]. Refer to Figure 3 to see the differences in the relative drag forces.

The second component of the viscous friction drag of the body is the pressure drag due to the boundary

layer. The boundary layer flow starts at the nose of the vehicle at the stagnation point, where the flow

velocity is equal to zero and the pressure is at its maximum.

Fluid velocity will increase as it flows over the body and the pressure will decrease. The pressure will

reach a minimum at the thickest section of the body. At this point the pressure will increase again and aim

for the stagnation value, but will not be able to achieve this value as the boundary continuously thickens

around the body. Therefore, full pressure recovery will never be reached. This creates a small wake behind

the vehicle, having a similar effect as the wake caused by the bluff body, albeit not as large. Refer to Figure

4 [10].

Figure 4 - Wake produced by different body shapes [10]

The sum of the pressure drag and the viscous friction drag is called the profile drag and is largely based on

the wetted surface area of the body of the vehicle.

Induced Drag

The induced drag of the body is a result of the lift generated by the airflow around the vehicle [11]. Figure

5 can be used to describe the induced drag on a wing [12]. The lift force increases or decreases as the

airflow around the airfoil changes. As the angle of attack of the airfoil is increased (by tilting the nose

upwards), the fluid experiences acceleration over the upper surface of the airfoil. This acceleration reduces

the pressure on the same surface. This also means that the airflow along the underside of the airfoil will be

slower, increasing the pressure. This increase in pressure as well as the change in direction of the flow,

exerts a force on the airfoil. This force is called the lift force. The increase in the lift force also means that
9
the resultant force (perpendicular to the chord line of the airfoil) between the lift force and the drag force is

increased. As the resultant force increases, the drag force will also increase. This drag force is referred to

as induced drag.

Figure 5 - Induced drag and lift [12]

This also applies to the streamlined body of a solar-powered vehicle. In this instance the angle of attack is

usually zero, but the lift generated is still upward. This results in the downward force on the tires being

reduced, albeit marginally. The lift induces vortices due to the pressure differences between the top and

bottom of the body. These vortices use kinetic energy and this increases the drag of the vehicle. The

vehicle is, however, not designed to generate lift, so induced drag is therefore not a major concern during

the design procedure.

Interference Drag

Interference drag is caused by local flow disruption at the intersections of different components of the body

[13]. These include the fairings-to-body and canopy-to-body junctions.

10
Table 2 - Relative sizes of the drag components and variables to which each is proportional [5]

Separation Interference
[%] Viscous Drag Induced Drag
Pressure Drag Drag
Holes, Seams,
Proportional to Frontal Area Wetted Area Lift Manufacturing
etc.

Medium to
Bluff Body Large Small Small to Large
Large

Approximately Approximately
Streamlined Body Large Small to Large
Zero Zero

Referring to Table 2 and to Figure 3, it is clear that the streamlined body has a much larger viscous drag,

but the pressure drag has a much larger effect on the total drag of a bluff body. It is also clear that the

separated flow caused by a bluff body is much larger and will have a larger suction force than the separated

flow caused by the streamlined body.

Solar-powered vehicles will therefore always be designed to be streamlined in order to minimise drag.

2.2 The Boundary Layer

The boundary layer is the layer of fluid flow adjacent to the body surface. Here the effect of viscosity is

important. Viscosity is the ability of the fluid to resist motion, thus the higher the viscosity of a fluid, the

larger the force required to displace the fluid.

The following paragraph, in conjunction with Figure 6 [14], describes the boundary layer:

Imagine that the body moves forward such that the air velocity relative to the body surface is U∞ m/s, called

the free stream velocity. The fluid next to the body surface does not move relative to the body surface, in

other words, it has zero velocity. This is called the no-slip condition. The layer adjacent to this zero

velocity layer, moves at a velocity relative to the surface of the body, but still much lower than U∞. The

velocities of the following layers will progressively increase until the speed is equal to U∞. These layers

that take the velocity from zero at the surface of the vehicle body to a value equal to the free stream velocity

U∞, are called the boundary layers.

11
The boundary layer becomes thicker as the flow progresses over the surface of the body. The thickening

of the boundary layer is described in [11] using Eq. 2. 4:

𝑣𝑥 2. 4
𝛿~√
𝑈∞

where the x represents the distance from the leading-edge and the v is the kinematic viscosity. δ refers to

the boundary layer thickness while U∞ is he free stream air velocity.

Figure 6 - Boundary layer along a thin flat plate [14]

The air flowing over the body is influenced to a greater extent by the body as the flow progresses further

down the length of the body. The top layers in the boundary layer are impeded by the stationary layer next

to the body surface, but still each layer will have a higher velocity than the layer beneath it. Eventually the

layers will reach the free stream velocity (U∞). This takes time (i.e. distance) meaning that the boundary

layer has to start out thin and build on itself as the flow progresses.

2.2.1 Laminar and Turbulent Flow

Laminar flow of fluids is characterised by airflow progressing in smooth sheets without any mixing, much

like the way light travels in straight beams [5]. Most flow patterns over objects start out as being laminar

but then transition to turbulent flow. The point where this occurs is called transition, which is largely

governed by the Reynolds Number [11].

12
As soon as this occurs, the fluid streams cross each other and this requires energy. The different layers

absorb energy from neighbouring streams, and these then result in the lower streams to gain more energy

and thus reach higher velocities. Referring to Figure 6 [14], it can be observed that the boundary layer

thickens due to the increased velocities. The turbulent fluid streams also absorb energy from the free stream

air, thus the turbulent boundary layer grows even thicker. Within the turbulent layer, there is still flow that

remains attached to the surface of the body, called the laminar sub-layer or viscous sub-layer. This layer

only accounts for about 1% of the turbulent flow, but the shear rates are much higher than the shear values

in the remaining turbulent layer [5]. This means that the laminar sub-layer accounts for most of the viscous

drag in the turbulent flow layer.

Natural Transition

The laminar boundary layer can become unstable at a certain critical Reynolds number, with the growth of

viscous instability waves (Tollmein-Schlichting waves). This generally occurs when the free stream

turbulence level is low, <1% [15]. At this stage, the waves are regarded as 2-D, but as they grow they

become increasingly non-linear.

13
Figure 7 - Natural transition process [15]

When this happens the flow becomes predominantly 3-D and results in turbulent spots. This is the point

where transition occurs. These spots will grow and mix with the laminar layer until the layer is fully

turbulent (see Figure 7 [15]).

Bypass Transition

When free stream turbulence levels are high (>1%), regions 1 through 3 in Figure 7 [15], which can be

referred to as the linear stages; is bypassed. This means that the turbulent spots are formed directly in the

boundary layer by the disturbances from the free stream flow [15]. Bypass transition is, however, not only

caused by free stream disturbances. It can also be caused by the effect of the surface roughness; thus

transition occurs due to disturbances at the wall of the geometry.

2.2.2 Separation

As mentioned in section 2.2.1., flow separation is the phenomenon when the fluid flow can no longer stay

attached to the body, and thus transitions from laminar to turbulent flow.

To the rear of the body, as the flow progresses downwards, the pressure becomes greater (adverse pressure

gradient). The pressure gradient will continue to increase and become steeper and the fluid will flow against

the pressure gradient as far as possible. The fluid, however, does not have enough energy to flow all the
14
way to the top of this pressure hill and the boundary layer will eventually separate from the surface [16].

At this point, flow can be pushed back near the surface. Refer to Figure 8. At this separation point, the

velocity gradient at the surface will become zero. It can be realised that at this point there is a stream line

that leaves the body surface, dividing the forward and reverse flow. This is the line of zero velocity and

the point where it leaves the body surface is called the point of separation.

Figure 8 – Schematic of flow separation at a wall [17]

Turbulent flow layers can support much steeper positive pressure gradients without experiencing flow

separation. The separation point is moved backwards when the flow is turbulent due to the energizing

effect of turbulent flow, where the kinetic energy levels of adjacent layers, such as the fluid layers close to

the body wall, is increased [18].

Fully separated flow results in vortices behind the body, consuming extra energy. Separation can also result

in air being pushed back, eventually moving faster than the body and also consuming extra energy.

2.3 Transition Modelling with CFD

Simulating flow around bodies using CFD, is gaining popularity in modern day research studies. Reasons

for this include the elimination of the use of wind-tunnels, which are expensive to operate; and by using

CFD, more iterations of the study can be completed in a smaller time frame (provided the computational

power is sufficient). There are many reliable turbulence models, allowing for accurate simulation results

to be obtained for turbulence modelling [19]. While simulating turbulence is important, an even more

important factor is the transition prediction of laminar flow on bodies.

15
Several simulation methods are available to compute the transition over a body, each with its own

limitations. All of these methods solve the Navier-Stokes equations, but with differences. The first of these

is the Direct Numerical Simulation (DNS) method. DNS solves the full unsteady-state Navier-Stokes

equations [15]. This method does not require a turbulence model for turbulence closure as no Reynolds

averaging is used, but in order to capture the small scales of turbulence, a very fine mesh is required; thus

large computational power. Even though great advances in computational technology have been made, it

is still not feasible to use the DNS method to compute transitional flow across a 3-D model.

The second method is the Large Eddy Simulation (LES) method. This method works on the same principle

as DNS, which is to solve the Navier-Stokes equations, but on a much coarser mesh. This means that only

the large scale eddies are directly resolved, while small scale eddies are modelled using an eddy viscosity

assumption. The small scale eddies are generally more isotropic; thus simpler, universal models can be

used rather than standard Reynolds stress models [20]. The LES method therefore computes the large scale

eddies using a fine grid and the small scale eddies are computed using sub-grid models.

Reynolds Averaged Navier-Stokes (RANS) modelling is the third commonly used method to model the

transition. This is a useful method from a design perspective, as a reasonable compromise is made between

accuracy and expense [21]. This means that the method is more cost effective than previous models but

will lower accuracy, though not to such an extent that the results are not reliable.

The most recent model is a correlation-based transition model called the Local Correlation-based Transition

Model (LCTM) [19]. This model uses only local variables and gradients, as well as the wall distance [22].

The intermittency function developed for this model is coupled with the SST k-ω turbulence model, turning

on the production term for the turbulent kinetic energy behind the transition location.

One realization of this LCTM concept is the γ-Reθ transition model.

2.3.1 The γ-Reθ Transition Model

Langtry described a model for transition prediction (Langtry et al. [23], Menter et al. [24], Menter et al.

[25] and Langtry et al. [26]) that is based on two transport equations, using only local information. The

first of these two transport equations is the intermittency, γ, which can be described as the fraction of time

16
during which the flow is turbulent, measured at a specific point in the transition region. The momentum

thickness Reynolds number is not used to trigger the onset of transition, but rather the strain-rate Reynolds

number:

𝜌𝑦 2 𝛿𝑢 𝜌𝑦 2 2. 5
𝑅𝑒𝑣 = = 𝑆
𝜇 𝛿𝑦 𝜇

where y is the distance from the wall, ρ is the density, μ is the molecular viscosity and S is the absolute

value of the strain-rate. What is of importance with regard to Rev is the relationship between its maximum

value inside the boundary layer and the momentum thickness Reynolds number (Reθ). As the boundary

layer grows, y2S will also increase until the critical value for Rev is reached.

The intermittency transport equation can be summarised as in Eq. 2.6 [23]:

𝛿(𝜌𝛾) 𝛿(𝜌𝑈𝑗 𝛾) 𝛿 𝜇𝑡 𝛿𝛾
+ = 𝑃𝛾1 − 𝐸𝛾1 + 𝑃𝛾2 − 𝐸𝛾2 + [(𝜇 + ) ] 2. 6
𝛿𝑡 𝛾𝑥𝑗 𝛿𝑥𝑗 𝜎𝑓 𝛿𝑥𝑗

Where the transition sources are defined as:

𝑃𝛾1 = 𝐹𝑙𝑒𝑛𝑔𝑡ℎ 𝜌𝑆[𝛾𝐹𝑜𝑛𝑠𝑒𝑡 ]𝑐𝑎1 2. 7

𝐸𝛾1 = 𝑐𝑒1 𝑃𝛾1 𝛾 2. 8

Flength is a correlation used to control the transition area length, while the onset of transition (Fonset) is

controlled by the strain-rate Reynolds number, Rev.

This model is usually coupled with the SST k-ω turbulence model (Shear Stress Transport model). Many

correlation-based transition models work on the principle of momentum thickness, which is strictly a 2-D-

concept [19]. However, the γ-Reθ transition model avoids this limitation and is able to compute transition

in 3-D flow models.

17
To couple the transition model with the SST k-ω turbulence model, the following equations are used:

𝛿 𝛿 𝛿 𝛿𝑘
(𝜌𝑢𝑗 𝑘) = 𝑃̃𝑘 − 𝐷
̃𝑘 + 2. 9
(𝜌𝑘) + ((𝜇 + 𝜎𝑘 𝜇𝑡 ) )
𝛿𝑡 𝛿𝑥𝑗 𝛿𝑥𝑗 𝛿𝑥𝑗

𝛿 𝛿 𝑃𝑘 𝛿 𝛿𝜔 2. 10
(𝜌𝜔) + (𝜌𝑢𝑗 𝜔) = 𝛼 − 𝐷𝜔 + 𝐶𝑑𝜔 + ((𝜇 + 𝜎𝑘 𝜇𝑡 ) )
𝛿𝑡 𝛿𝑥𝑗 𝑣𝑡 𝛿𝑥𝑗 𝛿𝑥𝑗

𝑃̃𝑘 = 𝛾𝑒𝑓𝑓 𝑃𝑘 2. 11

̃𝑘 = 𝑚𝑖𝑛(𝑚𝑎𝑥(𝛾𝑒𝑓𝑓 , 0.1), 1.0)𝐷𝑘


𝐷 2. 12

Where Pk and Dk are the original production and destruction terms for the SST model and γ eff is the

maximum value between γ and the modified equation for separation induced transition, γsep.

By coupling the intermittency, γ, with the above-mentioned turbulence model (SST k-ω turbulence model),

the production term is turned on. Regarding the modelling of transition, one can capture the effect of the

large free-stream turbulence levels on the laminar boundary layer, as well as the related increase in the skin

friction and heat transfer. In order to capture the non-local influence of the turbulence intensity, it is

necessary to solve for the transition onset momentum-thickness Reynolds number (Reθ). The turbulence

intensity changes due to the decay of the turbulent kinetic energy in the free stream layer, as well as changes

to the free stream velocity outside the boundary layer. The transition model therefore solves for the

intermittency, γ, as well as the transitional momentum-thickness Reynolds number, Reθ, and is thus named

the γ- Reθ transition model.

According to Menter et al. [19], the intermittency equation has been formulated to take into account the

rapid onset of transition during the separation of flow and can also be formulated for low free stream

turbulence intensity flows and cross-flow instabilities. These modifications make the transition model

extremely compatible with modern CFD codes such as STAR-CCM+.

2.4 CFD as a Design Tool

The external aerodynamic simulation process for simulating the flow around road vehicles, is described by

Mansor et al. [27] as having four important steps. These four steps are the model geometry setup, mesh
18
generation, numerical iterations and post processing. Yakkundi et al. [28] also indicates that the CAD

model is generated using a suitable software package, after which it undergoes surface repair during pre-

processing. This relates to Mansor’s model geometry setup.

Both Mansor and Yakkundi indicate the next step as meshing of the model or mesh generation. This is the

process of dividing the volume of the model into smaller elements in order to be able to solve the flow field

around the vehicle. Refinements is incorporated in specific areas around the model where larger elements

would not be able to capture the geometry effectively.

A mesh independence study is an important step to conduct. This ensures that the solution obtained during

the simulations are independent of the mesh resolution. In other words, reducing the mesh size further, has

no effect on the result. If one were to further refine the mesh after mesh independence has been reached,

computation time will be increased, and the result will remain within a very small margin of the previous

simulation. During this same step the different turbulence models can be compared in order to test the

accuracy of the different turbulence models. The most common turbulence models used in CFD simulations

are the k-ε and the k-ω turbulence models. Both Mansor et al. [27] and Yakkundi et al. [28] found the k-ω

SST turbulence model to be more accurate.

Boundary layer conditions are of utmost importance when simulating flow around a road vehicle. The fluid

domain should be defined as air. The boundary in front of the vehicle is defined as the inlet of the domain

and should be defined in the CFD simulation as a velocity-inlet with the same velocity as the speed at which

the vehicle is being simulated to move forward. The outlet of the domain is set as a pressure-outlet with

the pressure at 0 Pa (relative to atmosphere). The inlet domain should be set at least two vehicle lengths

ahead of the vehicle to ensure that the flow is fully developed (and laminar) by the time it reaches the

vehicle body. The outlet domain is set at least three vehicle lengths behind the vehicle. This ensures that

the separation of flow and wake generation is properly captured. The vehicle surface should be treated as

a no-slip condition while the walls of the domain can be treated as free slip conditions as well as symmetry

walls.

The last step in this basic methodology is to analyse the results of the simulations. The simulations have to

be run until a converged solution is achieved, usually by first solving for only turbulence after which the

19
transition model can be activated. This enables the solution to diverge easier by removing some of the

variables to be solved for at the beginning of the simulation. The results are evaluated in the step referred

to as post-processing and is usually compared to the experimental results (if available) at this stage.

2.5 Aerodynamic Design

For almost all vehicles, aerodynamics play an important role, whether it is to increase fuel economy as in

the case of production vehicles or to increase down force in race vehicles. Solar-powered vehicles often

run on as much energy as used by a hairdryer, so in this case reducing aerodynamic drag to the absolute

minimum is essential.

When considering the purpose of the vehicle, this need is further enhanced and the essentiality is

highlighted. The solar-powered vehicle is built to compete against other solar-powered vehicles. All of

these vehicles are only powered by energy obtained from the sun and each team will strive to reduce the

energy consumption of their vehicles to a minimum; in most cases by reducing the aerodynamic drag. If

this is not done, the vehicle will have a clear disadvantage. By reducing the aerodynamic drag and thereby

improving the energy consumption of the vehicle, the vehicle will be able to travel further at higher speeds,

and therefore have a better chance at winning the competition.

The following sections will describe how the vehicle design is affected by the aerodynamic forces on the

body and how the shape of the body can be altered to reduce these aerodynamic effects. During the design

process it is always important to reduce the frontal area as this has a direct effect on the drag force. A

person holding his or her hand out a car window when the car is moving at a high speed, will feel a

noticeable difference between holding their hand vertical (palms perpendicular to the moving air) than when

their hand is rotated to be horizontal to the ground. This movement reduces the frontal area and in effect

the aerodynamic drag force.

2.5.1 Vehicle Body Shape

The 3-D shape with the least amount of aerodynamic drag is that of a streamlined body [16], such as a water

droplet, but due to the requirement of a large flat surface to accommodate the solar array, this droplet shape

20
is not suited for the body shape of a solar-powered vehicle. So the droplet shape is modified to have a flat

tail section, to accommodate the solar array.

In the case of a droplet shape, the drag is at a minimum when in free stream air [29]. Bodies travelling in

close proximity to the ground, such as solar-powered vehicles, experience a venturi effect as the air flows

between the body and the ground plane [5].

Figure 9 - Summary of how camber effects the lift force on the body in freestream and in close ground

proximity [5]

The ground effect increases the downforce on the droplet shape as the stream lines around the shape become

compressed between the body and the ground as is illustrated in Figure 9. The streamlines concentrate

more on the underside of the body than on the top, and this venturi effect produces downforce. Cambering

the shape changes the flow patterns around the body, reducing the venturi effect and therefore the

downforce on the body.

When the cambered body is in freestream, the streamlines experience greater acceleration over the top of

the body. This increase in velocity reduces the pressure on top of the body, resulting in lift. When a

cambered body is in close proximity to the ground, this lift can be eliminated, increasing the stability of the

vehicle.

Another method to reduce the drag is to change the top view of the body shape. In the case of a rectangular

shape, the solar array would fit easier onto a body of constant width, making it more desirable [29]. Even

though a solar array would fit onto the rectangular shape with more ease, a rectangular body shape is not

the most aerodynamic. However, the drag can be reduced by tapering the trailing edge of the body. This

21
complicates the solar array design as the curvature of the body makes it difficult to fit a rectangular array

without wasting area on top of the body.

Figure 10 - Tapering of the trail edge of a bluff body (boat-tailing) [30]

Referring to Figure 10, the effect of tapering the trailing edge of a bluff body can be observed. This is

commonly called “boat-tailing” [30]. Even though this is based on drag reduction of bluff bodies, the

theory can be applied to the shape of the solar vehicle body as well. By tapering the body at 22°, the drag

can be reduced by more than 10%. In the case of solar vehicle bodies, it is not always possible to taper the

vehicle to that extent, as the competition regulations limit the length of the vehicle and space is still required

to fit 6 m2 of solar cells.

The static stability of the main body is also important, as it will affect the dynamic stability of the entire

vehicle. According to McCormick [31] static stability refers to the tendency of an airplane to return to a

trimmed condition when disturbed under steady conditions. It does not refer to any motion it undergoes

after the initial disturbance. This theory may also be applied to the main body of the vehicle as it is

essentially a wing of an airplane. For the vehicle to be in equilibrium, the forces and moments acting on

the vehicle must be in a direction to constantly force the vehicle back to its equilibrium position. In other

words, the forces acting on the body must balance around the centre of gravity of the vehicle. This point

where the forces are in balance, is referred to as the trim point.

22
Figure 11 - Body at trim [32]

Referring to Figure 11 [32], the body will be in balance or trim when the moment about the centre of gravity

is zero. This is where the lift coefficient is equal to 1 and the angle of attack is 0°. The lift coefficient must

be positive at 0 angle of attack in order to generate lift (in the case of an airplane). If the angle of attack

(α) is increased, in other words, an increase in the lift coefficient (CL), and the moment increases about the

centre of gravity, the body will become unstable. The opposite will happen if the moment decreases with

increasing CL. Therefore, for the body to be stable, two conditions must be met:

𝑑𝑀
<0 2. 13
𝑑𝛼

This means that the moment about the centre of gravity of the body must be negative. The other condition

that has to be met for vehicle stability is that the centre of gravity must be ahead of the centre of lift. See

Figure 12 [33].

Figure 12 - Airfoil contribution to longitudinal stability [33]

If the centre of gravity is behind the centre of lift, the body will experience a positive moment. When the

lift force increases with increased α, the moment about the centre of gravity will increase linearly and will

23
cause the body to become unstable. What eventually will happen is the front wheels of the vehicle will lose

traction on the road surface, which can result in an accident. If the centre of gravity is in front of the centre

of lift, the body will have a negative moment. This will have an effect of pushing the nose of the body

down in the case of increased α. The vehicle’s front wheels will thus remain in contact with the road surface

and thus stable on the road. This can also be defined in terms of the neutral point:

𝐶𝑀𝛼 = (ℎ − ℎ𝑛 )𝐶𝐿𝛼 2. 14

This equation states that the slope of the momentum curve, CMα, is equal to the slope of the lift curve, CLα,

multiplied by the dimensionless distance of the centre of gravity behind the neutral point of the body, (h-

hn). However, for the body to be statically stable the momentum must be negative, meaning the centre of

gravity must be ahead of the neutral point. The value (h – hn) is the static margin and must be at least 5%

of the mean aerodynamic chord of the body [31].

2.5.2 Nose design

An effective way to delay the transition of flow from laminar to turbulent is to design the nose of the body

to have a favourable pressure gradient. Imagine a fluid molecule of mass m flowing over the nose of the

vehicle, which is usually convex in shape. This will result in the fluid molecule accelerating and this

resultant increase in velocity will result in a reduction in pressure [5].

2
This molecule experiences a centrifugal force (𝑚𝑉 ⁄𝑅 ), pushing it away from the surface of the body. V

is the local streamline velocity and R is the local radius of curvature of the body. In order for the molecule

to remain attached to the body, the local pressure gradient must be such as to push the molecule against the

body. In other words, the centrifugal force acting on the molecule must be balanced by the pressure gradient

[34]. This means that the free stream air pressure must be greater than the local pressure at the surface of

the body. By increasing the centrifugal force, by either increasing the free stream velocity or by decreasing

the local radius of curvature, the local pressure will decrease. The resulting effect will be that the free

stream pressure will be greater than the local pressure and the molecule will remain attached to the body;

thus the flow would remain attached. This attached flow could be laminar or turbulent depending on

whether transition has been reached.


24
A negative pressure gradient results during acceleration, as the pressure at one point is lower than the point

directly before it. This negative pressure gradient has the effect of stabilizing the laminar boundary layer,

and is thus favourable.

The pressure coefficient (or coefficient of pressure), Cp can be defined by Eq. 2. 15:

𝑃𝑙𝑜𝑐 − 𝑃∞
𝐶𝑝 = 1
2. 15
𝜌𝑉∞2
2

According to Bernoulli’s Principle, Cp can also be expressed by Eq.2. 16:

Vloc 2 2. 16
Cp = 1 − ( )
V∞

Here Ploc and Vloc are, respectively, the local pressure and velocity values near the body surface. Looking

at Eq. 2. 15 it is clear that the value for Cp is a measure of the local pressure. It is assumed that the pressure

across the thickness of the boundary layer is constant, as the boundary layer is very thin. The pressure is

equal to the pressure just outside of the boundary layer, P∞ of Eq. 2. 15. Therefore, if the value for Cp is

known, the rise and fall of the surface pressures can be calculated. By tracking the slope of the Cp curve

along the body, the pressure gradient curve can be determined.

Eq. 2. 16 indicates that the value for Cp must be equal to 1 when Vloc is equal to zero. This point is termed

the stagnation point. The air accelerates over the nose of the body and this causes a rapid reduction in the

pressure at the same point. Nearing the thickest part of the body, the flow will continue to accelerate, but

at a slower rate. It results in further reductions in the pressure and the Cp value becomes negative. At the

point of maximum thickness and minimum pressure, the pressure will start to increase again. This is called

the pressure recovery region, where the Cp value progresses in the positive direction [5]. It is usually in the

pressure recovery region that the flow separates from the body.

2.5.3 Interference Drag Reduction

As described in Section 2.1, interference drag is the result of irregularities on the surface body such as

manufacturing defects, ventilation scoops, poorly fitted body panels, wheel arches, etc. This section

describes design procedures to reduce the effect of interference drag.

25
Junction Flows

Junction flows refer to the junctions between the main body of the vehicle and appendages such as the

wheel fairings and the canopy for the driver. The drag increase ∆Dj due to these junctions [5] can be

approximated by the following equation:

𝑡 𝑐 2
∆Dj = 𝑞𝑡 2 [0.75 ( ) − 0.0003 ( ) ] 2. 17
𝑐 𝑡

Where q is the dynamic pressure 0.5ρV2, c is the chord length and t is the thickness of the airfoil-shaped

appendage. Referring to Figure 13 [35], the chord length and the thickness of the appendage, relating to

the above equation, is clearly shown.

Figure 13 - Junction flows between the fairing and the body [35]

The junction flow is described by Boermans et al. in [36]. The flow ahead of the appendage is laminar, but

this laminar boundary layer cannot overcome the adverse pressure gradient created by the junction between

the body of the vehicle and the appendage, or the body-appendage junction. This adverse pressure gradient

causes the laminar boundary layer to transition. The turbulent boundary layer is, in turn, forced to separate

from the surface of the body due to this adverse pressure gradient. The adverse pressure gradient is formed

due to the junction and the flow will roll up into a vortex at the root of the appendage, as is seen in Figure

26
13. These vortices are of the rotational sense of the approaching boundary layer and because of this, the

vortices entrain higher speed fluid along the appendage which increases the drag in the junction region [37].

The simplest method to reduce the drag caused by junctions is to minimize the number of junctions. By

aligning the fairings with the outside lateral edge of the main body, the number of junctions can be reduced,

as well as the interference drag. Once these junctions have been reduced or eliminated, it is necessary to

minimize the drag of the body. The key here is to eliminate flow separation.

Inclination Angle of the Wheel Fairings

According to Tamai [5] the inclination angle (θf), as indicated in Figure 14 [5], of the wheel fairing to the

vertical plane should be as small as possible. Therefore, the leading-edge angle should rather be near

vertical, as a fairing that is near perpendicular to the body has a Cd (drag coefficient) value of about 0.08,

provided the thickness-to-chord ratio is about 0.43.

Figure 14 - Inclination angle of an appendage [5]

The Cd value increases minimally for inclination angles up to 10°. For every degree that the inclination

angle exceeds the 10° mark, the drag increases linearly by about 2.5% per degree of inclination. This can

be attributed to premature separation of flow, which is what should be avoided. This means that when

designing the fairings, it should be done so that the fairings intersect the body perpendicularly. Simpson,

however states that transition can be delayed by increasing the inclination angle of the fairing as it shifts

the vortex downstream [37]. This is based on experiments conducted by Ahmed and Khan [38]. By

27
designing a faring that joins the body perpendicular and sweeps back at an angle of less than 10°, it might

be possible to reduce the effect of the vortex.

Fillet Radius along the Appendage

Fillets can help to reduce the interference between the body and the appendage [39]. This is achieved

because the fillet eliminates or minimizes the effect of vortices that forms at a sharp intersection between

an appendage and the body of the vehicle. By implementing fillets where the appendage meets the body,

as indicated in Figure 15 [5], a drag reduction of as much as 20% can be experienced. The optimum fillet

radius, as indicated by the above-mentioned author, is 4-6% of the chord length of the appendage. This

value was originally experimentally determined by Hoerner [40].

Figure 15 - Fillet radius along an appendage [5]

It was also determined that a smaller fillet radius proved to be more favourable. This was experimentally

determined by Maughmer et al. [41], when wind tunnel tests revealed that a 10% radius on a sailplane had

a measurable reduction in the drag force compared to sailplanes with 35% and 45% radii.

Leading-edge Junction Geometry

If the leading-edge of the fairing is blunt, in other words, with no fillet (Figure 13 [35]), creating a sharp

transition from the fairing to the body, the adverse pressure gradient ahead of the fairing will cause the flow

to separate and form a vortex. This vortex, called a horseshoe vortex, will continue along the entire length
28
of the fairing. This will increase the shear stress along the fairing junction. The increase in pressure drag

beyond the trailing edge of the fairing is usually what increases the drag of the vehicle.

By implementing a leading-edge fillet of 63°, the separation of the flow can be eliminated. Using a fillet

with a 30° angle, separation can be greatly reduced [5]. Smoothing these leading-edge fillets further helps

to eliminate separation. Another method to reduce the drag, using a leading-edge fillet, is to implement a

linear fillet. The linear fillet showed as much as a 7% drag reduction over a parabolic fillet, according to

Maughmer et al. [41].

2.5.4 Canopy Drag Reduction

Bubble canopies are the most popular design and will be implemented in the design. These bubble canopies

should allow the driver sufficient space to move forward with clear peripheral visibility. The canopy should

also not be too large as this reduces the area available to the solar array. The points discussed in the previous

subsection should also be applied to the junction between the canopy and the body.

Moving the canopy as far backwards as possible is also essential to prevent premature flow separation.

Viewing the canopy from the front, it should be as narrow as possible in order to create a narrower

separation bubble around the canopy, thus more laminar flow on the body.

A major design variable for canopies is the length-to-height ratio, L/h. According to Tamai [5], it was

reported that certain university solar car teams claimed that their vehicles had canopies with little or no

separation. These canopies had L/h ratios of about 4.5, with the highest point of the canopy at about the

middle of the canopy. However, the author also states that a safe optimum L/h ratio for a canopy would be

6. This is based on data in Figure 16.

29
Figure 16 - The upper and lower bounds of the canopy drag as a function of the L/h ratio [5]

The drag values of canopies also change as the shape of the canopy is changed. For the above L/h ratio of

6 and a trailing section that tapers into a point, the approximated drag value for the canopy is 0.045 (based

on the frontal area of the canopy) [5]. It is quite clear by referring to Figure 17 [5] that designing a canopy

to have any type of blunt end will increase the drag.

30
Figure 17 - (a) Illustration of baseline canopy with L/h = 6. (b) Base line canopy with a blunt nose. (c) Baseline

canopy with a blunt tail. (d) Baseline canopy with a blunt nose and tail [5]

The L/h ratio method is an accurate design methodology to be used for the canopy design. But it should

also be noted that to allow sufficient space for the solar arrays while still allowing enough space for the

driver and roll bar, the L/h ratio of the canopy might be much closer to a value of 3 to 4. The shape of the

canopy in terms of leading and trailing edges should also be kept in consideration.

2.5.5 Trailing Edge Thickness of the Main Body

In many 2-D, as well as 3-D simulations, it is often assumed that the trailing edge of an airfoil is perfectly

sharp. However, this is not the case with solar vehicle bodies. The ideal would be to design for a knife

edge trailing edge, but this is not always possible.

31
Figure 18 - Trailing edge thickness plotted against the percentage increase in drag area [5]

This drag increase caused by the trailing edge, is estimated by the equation described by Hoerner [39] and

is based on the frontal area of the main body:

4
0.34 𝑡𝑡𝑒 3 2. 18
∆𝐶𝑑 = 3 ( )
√𝐶𝑑,𝑓𝑟𝑜𝑛𝑡𝑎𝑙 2𝑡

where 𝐶𝑑,𝑓𝑟𝑜𝑛𝑡𝑎𝑙 is the drag coefficient of the main body, based on its frontal area, 𝑡𝑡𝑒 is the thickness of

the trailing edge and t is the maximum thickness of the main body of the vehicle.

Referring to Figure 18, it is clear that a 5 mm thick trailing edge causes an increase of 1% in the drag value.

An increase of 1% is regarded as significant and therefore it is best if the trailing edge does not exceed 5

mm. It would be ideal if the trailing edge of the vehicle was a knife edge, but practicality should also be

considered as the vehicle has to have rear taillights, as per regulation.

2.6 Popular Solar-powered Vehicles

Over the years certain institutions have gained extensive experience in developing and manufacturing solar-

powered vehicles. These institutions have competed in various competitions, including the World Solar

Challenge held in Australia every second year [42]. The following teams are placed respectively first,

second and third globally, based on performances at the 2013 World Solar Challenge.
32
Delft University of Technology: Nuna7 (Nuon Solar Car Team)

The team’s car line-up is called Nuna. They have already produced six vehicles, having started in 2001.

The Nuna6 (displayed in Figure 19 [43]), which competed in the 2011 World Solar Challenge, achieved an

overall second position. The body of this vehicle is a sandwich construction, with face material being

carbon fibre and a foam core. “TeXtreme” Carbon fibre fabric was used, which is also used in the

construction of Formula 1 cars. The roll bar used was made of titanium and aramid reinforced parts were

used to protect the driver during possible collisions. The vehicle had an aerodynamic drag coefficient of

10% less than the previous model, the Nuna5.

Figure 19 - Nuna6 solar vehicle [43]

The Nuna7 (Figure 20 [44]), which is the university’s first four-wheel vehicle, was built to compete in the

World Solar Challenge in Australia in 2013 (Figure 20 [44]). Four wheels were required according to the

competition regulations; a four wheel vehicle being more stable than the three wheel designs. This vehicle

uses the same body shape as the Nuna6, but to cover the wheels two long fairings were used. The two

fairings are replicas of each other, wide enough for the driver to be seated between the front and rear wheels

on the left hand side of the vehicle. The space on the opposite side to the driver is used to install various

components such as the battery case, thus ensuring even weight distribution across the vehicle.

33
Figure 20 - Nuna7 solar vehicle [44]

The Nuon Solar Team is currently ranked first in the world for the Challenger Class of solar-powered

vehicles after they finished first during the World Solar Challenge in 2013.

Tokai University: Toray

As with the Nuna range of Nuon Solar Car Team, Tokai University has always focused their designs on

three-wheel concepts (see Figure 21 [45]). Tokai University, from Japan, has always been the main

competition for the Nuon Solar Car Team followed by Solar Team Twente. As mentioned, the Challenger

Class now regulates that vehicles must have four wheels, and Tokai University also built their first four-

wheel vehicle for the World Solar Challenge in 2013.

34
Figure 21 - Tokai University solar-powered vehicle [45]

This vehicle is different from the Nuna7 vehicle in that it has only one long fairing, also serving as the

driver cockpit. The remaining two wheels on the opposite side of the vehicle are covered by individual

fairings. Once again the body shape is that of a wing airfoil, thus enhancing the laminar flow on the body

of the vehicle.

Unlike the Nuon Solar Team’s solar-powered vehicle, the different components of this vehicle are mounted

in the main body, distributed evenly so that the centre of gravity is closer to the front of the vehicle. This

enhance the stability of the vehicle.

Solar Team Twente: The Red Engine

Another popular solar-powered vehicle, dubbed the “Red Engine”, was built by Solar Team Twente from

the Netherlands (see Figure 22 [42]). This vehicle has a different concept to that of the other teams already

discussed. The basic body shape is also that of an airfoil, but it is lofted towards the centre of the vehicle.

This more rounded shape allows the driver of the vehicle to be seated in the centre of the vehicle.

35
Figure 22 - Solar Team Twente's "The Red Engine" solar-powered vehicle [42]

This vehicle further has a single shorter fairing covering each of the four wheels. The two rear wheels’

track width is narrower than the front wheels. All the components are installed in the main body, again so

that the centre of gravity is towards the front end of the vehicle.

One can compare the effectiveness of these designs with one another. This enables the most promising

design features to be identified, which can then be used to enhance the aerodynamic design of the NWU’s

solar-powered vehicle. In section 2.4 it is indicated that the frontal area of the vehicle should be reduced

as it has a direct effect on the aerodynamic drag of the vehicle. Keeping this in mind while evaluating the

design of the three vehicles discussed, it makes sense to only look at the Nuna7 and the Toray. This is

because both these vehicles have the driver seated on the side of the vehicle. This greatly reduces the frontal

area of the vehicle and can therefore be regarded as important design feature. The fact that the Toray also

has two shorter fairings on the one side, means that the overall width of these fairings are narrower than the

single fairing of the Nuna7. This adds to the reduction of the frontal area. Furthermore, the friction drag

between the air and the surface of the vehicle also contributes to the drag of the vehicle. The two shorter

fairings covering the wheels opposite to the driver’s side of the Toray means that the vehicle has a smaller

wetted surface (body surface in contact with the air). Theoretically this should help to further reduce the

drag of the vehicle. The Red Engine has a larger wetted surface than any of the other two vehicles discussed

36
and has a much larger frontal area. Theoretically this means the aerodynamic drag of this vehicle would

be higher. All of the vehicles have one thing in common, namely smooth junctions between the different

body appendages. This is achieved by fillets to reduce the vortex effects that the body has on the air flow

past it and to extend the laminar flow.

The most promising features that can be carried forward to the design process is to move the driver to the

side for an asymmetric design and to reduce the width of all fairings to the absolute minimum that is

required for the proper functioning of the internal components. Effort should be made to reduce the wetted

surface of the vehicle in order to further reduce the aerodynamic drag of the vehicle. Properly sized fillets

should also be incorporated into the design of the junctions.

2.7 Summary

Aerodynamically, solar-powered vehicles can be improved by reducing or even eliminating the drag

components contributing to the drag force on the vehicle. These include the pressure drag, friction drag,

induced drag and interference drag. These components are directly affected by the shape of the vehicle and

the different design factors discussed. It is important to choose the values for these factors, such as the

body shape, nose and trailing edge design as well as canopy and fairing designs, as near the theoretical

optimum as possible. The most promising factors of previous designs is to move the driver to the side of

the vehicle, an asymmetric design, and to reduce the width of the fairings. This reduces the frontal area of

the vehicle and should therefore help to reduce the aerodynamic drag of the vehicle. One should further

try to reduce the wetted surface of the vehicle which can also contribute to the drag reduction. These

changes are incorporated into a methodology which describes four basic steps, namely model geometry

setup, mesh generation, numerical iterations and post-processing.

37
Chapter 3: Methodology

By researching previous methods and studies of a similar nature, a preliminary design of the vehicle can be

developed that will be closer to the final design, thereby reducing the design time considerably. The most

important aspects of this methodology will be covered in this section, these same aspects has already been

summarised by Caroll [29]. They are directly applicable to the design process of a solar-powered vehicle.

The methodology proposed in this section focuses specifically on the aerodynamics and the body design of

the vehicle.

The methodology is a generic method described by Caroll [29] and is based on published methods such as

those described by authors Abe and Starr [46] and Ullman [47]. The following six steps describe the design

process that will be followed when designing the body shape of the solar-powered vehicle. It is adapted

from the mentioned authors’ works and can relate to the process described by Kissinger [48]:

1. Review of competition regulations pertaining to the body design of the vehicle and of previous

works. Perform software validation.

2. Determine the initial design parameters to be used to optimise the body of the vehicle, namely

the frontal area and the drag coefficient.

3. Design of preliminary vehicle body shapes (concept designs) based on solar-powered vehicle

dynamics as discussed by Caroll [29] and Tamai [5]. Here the model geometry is transferred

to a suitable CAD software package.

4. Analyse concept designs in terms of the drag coefficient as well as manufacturability. Select

a concept design. This step can be conducted using CFD simulations, implicating the mesh

generation and numerical iteration steps discussed in the literature survey would have to be

applied.

5. CFD analysis (based on the steps mentioned above) of the preliminary body design of the

vehicle will be conducted and models are compared during the post-processing stage. The

comparisons are based on CD and transition locations on the turbulent kinetic energy plots.

6. Revise the design by adjusting factors as described by authors mentioned in step 3 and results

found in step 5. This is also part of the post-processing step as described in the literature.
38
3.1 Review Regulations and Evaluate Existing Solutions

The first step is to review all of the competition regulations that will have a direct effect on the body design

of the vehicle. This means studying the competition regulations to ensure that the body design will conform

to all regulations, making it eligible to compete in the competition. Therefore, a good understanding should

be developed in order to have a good idea of what is regarded as being competitive. Existing solar vehicle

designs should also be studied in order to determine reliable designs.

3.1.1 Summary of previous work

The most common design for solar-powered vehicles is a vehicle with only three wheels – commonly with

two steering wheels in the front and only one driving hub motor in the rear. The most successful teams are

the Nuon Solar Team from the Netherlands, Tokai University from Japan and Solar Team Twente also from

the Netherlands.

These teams have many years’ experience in building solar-powered vehicles, resulting in vehicles that

have coefficients of drag values claimed to be below 0.1. The drag resistance caused by the vehicle’s body

is extremely low and thus little energy is required to overcome the resistance. More solar energy received

from the sun can therefore be used to overcome the rolling resistance (also a large factor to be considered)

and more power is available to maintain a higher average speed for a longer distance.

As indicated, the common design for a SV is based on three wheels. New regulations state that SVs should

have four wheels in order to compete in the main classes. The reason for this change is that the three wheel

vehicles are too unstable at high speeds. The organisations (Sasol Solar Challenge and World Solar

Challenge) also wish to encourage the technology to advance towards production line vehicles and not to

only be focused on designing newer and better race vehicles. The design for three wheel vehicles has been

optimised, so the current challenge is to design a four-wheel vehicle with the same drag coefficient figures

as the current three wheel designs.

3.2 Geometric Specifications and Initial Design Parameters

This design process is applicable to the body shape of the solar-powered vehicle, but other specifications

of the vehicle have a direct influence on the eventual shape of the vehicle. These specifications include the
39
space required for the solar array, the battery capacity and the shape of the battery packs. The type of motors

and suspension to be used will determine the width of the fairings covering the wheels and the shape of the

canopy is determined by the space requirements for the driver.

3.2.1 Solar Array and Body Shape

The maximum width of the vehicle should not exceed 1.8 m and the vehicle may not exceed 4.5 m in length.

The maximum height of the vehicle is 1.9 m. This leaves relatively limited space for the solar array; with

a maximum allowable area of 6 m2 (silicone mono-crystalline solar cells). The largest part of the body

should therefore be relatively flat or gently curved (such as seen in an airfoil of a wing) surface. This will

form the base onto which the solar cells are mounted.

3.2.2 Battery Pack

The size of the battery pack is limited according to weight and the maximum weight is 21 kg. The battery

pack is usually mounted in the nose of the vehicle in order to enhance the aerodynamic stability of the

vehicle. The battery pack is therefore limited by height, rather being longer and wider. If the battery pack

is mounted to the centre of the vehicle, the centre of gravity of the vehicle is shifted to the rear, which can

decrease the stability of the vehicle. If the battery pack is too high, it might require an increase in the height

of the main body section, increasing the frontal area of the vehicle.

3.2.3 Electric Motors and Wheel Size

Two popular drivetrains are used. The simplest drivetrain used is an electric motor mounted, in the vehicle

in such a position to give power to driving the wheels via a chain-link or a similar device. The other type

of motor is an in-hub electric motor. The electric motor is mounted inside the rim of the driving wheels.

These motors are usually at least 150 mm in width, but the more expensive models can be as narrow as 75

– 95 mm. The wider the motor, the wider the wheel fairings will have to be.

3.2.4 Suspension, Steering Mechanism and Wheel Size

Adding to the electric motor thickness, is the suspension. Numerous suspension designs can be

incorporated, each design with its own share of advantages and disadvantages. A commonly used double

wishbone suspension might require a fairing to be much wider and therefore will result in an increase in

40
the frontal area of the vehicle. The larger the wheel, the more space will be required at the same steering

angle as a smaller wheel. This, again, increases the frontal area due to wider fairings. These factors should

be given precedence when designing the body shape of the initial design.

3.2.5 Driver Position

The driver should be seated comfortably and have ease of access to all controls at any time. The driver

should be able to move forward 45° and 25° backwards, without his/her head touching the canopy shield.

This might result in a longer canopy being designed, which increases the wetted surface area of the vehicle.

He/she should also have a clear field of vision around him/her and his/her eyes should be no less than 700

mm above ground level. Furthermore, the vehicle must have 4 wheels and a ground clearance of no less

than 90 mm.

The initial design parameters for the frontal area and the drag coefficient were determined by reviewing

existing designs of solar-powered vehicles. This was done in order to design a vehicle that would be on

par with world leaders in the trade and to ensure that the vehicle would be able to compete with these other

solar-powered vehicles. A value for the frontal area of the vehicle of not more than 1 m2 was chosen, as

well as a value of 0.1 for the drag coefficient.

3.3 Create Concept Designs

As has been mentioned, previous solar-powered vehicles of other institutions needed to be studied. This

will give insight in the most popular designs and different aerodynamic designs that have been proven in a

similar competitions to the Sasol Solar Challenge. These should be used as a baseline from which to design

different concepts for the vehicle body shape, while constantly keeping in mind the geometric specification

mentioned in Section 3.2 of this chapter.

A number of different conceptual designs were evaluated and compared with each other in order to find the

design with the most favourable aerodynamic properties. The main areas focused on during the conceptual

designs, were the fairing configurations and the position of the driver, while still giving precedence to the

reduction of the frontal area and minimizing the effects of the interference and profile drag.

41
3.4 Selection of Best Concept Design

The drag force on the bodies of the different concept ideas, will be determined according to the simulation

model developed during the validation procedure. The results of these simulations should then be presented

to the whole team. During this step, the designer should point out the advantages as well as the

disadvantages of the different concept designs. The team will then have to discuss the pros and cons of

each of the designs and a consensus should be reached as to which design is best in terms of aerodynamics,

manufacturability and overall appearance.

3.5 Detailed Design

This step can only proceed once the concept design has been chosen. This is an iterative process and the

different aspects of the body will have to be adjusted in order to find the optimum shape with the best

aerodynamic properties. The development of the detail design is discussed in Chapter 7 of this study, once

the simulation procedure has been validated in the following chapters.

A preliminary body design for the solar-powered vehicle is developed from the selected concept design.

This design has to conform to the regulations discussed and the chosen design parameters. By referring to

the literature of Caroll [29] and Tamai [5], CAD parts of the different components of the vehicle, such as

the fairings and canopy, can be made. The main body section, onto which the solar array is integrated, is

chosen as the basis of the design. Around this part, the other components will be designed and integrated

according to the above-mentioned literature.

During this process the minimisation of the frontal area of the vehicle should constantly be given

precedence, while conforming to the regulations mentioned. Factors that will increase the frontal area of

the vehicle include the width and height of the canopy in order to allow enough headroom for the driver;

and space for a roll bar to protect the driver in the event of a collision. The front wheels are used to steer

the vehicle and there should thus be enough space inside the fairings to allow the wheels to turn without

the suspension and brake mechanism touching and possibly damaging the inner walls of the fairings. The

body must also be smoothed and sharp corners must be avoided to reduce the effects of interference drag.

42
These sharp corners can cause vortices to form which will increase the pressure drag of the vehicle.

Therefore, all edges are smoothed by the use of fillets.

43
Chapter 4: 2-D Validation: NACA 0018

Validation is the process of determining the degree to which a model is an accurate representation of the

real world [49]. This is one of the most commonly used definitions for validation and can be broken down

into several parts. Validation is a process and can thus not be conducted in only one step. The process also

requires that the accuracy should be determined, meaning that error and uncertainty will be factors [50].

When a model has been validated, it is implicated that the validation process is reliable and can be applied

to similar models under the same conditions. It can thus be used on other models to make accurate

assumptions of reality. It also gives assurance that the model has been tested against well-defined criteria,

such as experimental results obtained from wind tunnels, for example.

The 2-D validation process was therefore conducted in order to gain confidence in whether the transition

model would iterate correctly at low Reynolds numbers. This then meant that the transition model could

be applied in more complicated models, such as 3-D models. The transition model used in the validation

of flow around a 2-D airfoil was the γ-Reθ transition model, coupled with the k-ω SST turbulence model.

CD-adapco’s STAR-CCM+ was used as the solver.

4.1 Validation Setup

The geometry used was the NACA 0018 airfoil profile. The coordinates for this airfoil were obtained from

Profili 2.0 software [51]. This airfoil was used because of the availability of wind tunnel experimental

results as described by Timmer [52].

The coordinates of the airfoil were used to create a solid part in SolidWorks. A solid block, representing

the wind tunnel, was also created here, after which the airfoil volume was subtracted from the wind tunnel

volume. The fluid volume in the wind tunnel was thus represented by a solid part. The test domain for the

validation procedure was selected near the experimental values that have been described by Timmer [52].

The domain size was 2.6 m in length, 1.8 m wide and 1.25 m high, with an octagonal face as shown in

Figure 23. The wing model had a chord length of 0.25 m.

44
Pressure Outlet
Boundary Wall

Model
Velocity Inlet

Figure 23 - Domain with boundaries defined

The inlet boundary wall is treated as a velocity inlet, with the velocity set to 40 m/s. This will then represent

the speed at which the air flows past the model, with the model standing still relative to the ground. This

was also the method used by Mansor et al. [27]. For the model chord length of 0.25 m, this velocity setting

relates to the Reynolds number of 0.7x106 that was used in the experimental setup. The outlet boundary is

set to a pressure outlet with the pressure setting at 0 Pa. Both these settings have been discussed in the

literature. The boundary walls are treated as solid walls, and not as symmetry planes, because the test setup

was conducted in a closed domain with walls exhibiting a no-slip condition.

Due to the symmetric nature of the problem, only half of the geometry was used to generate the 3-D mesh;

in other words, half of the fluid volume. The geometry was divided into two regions to enable the use of

an overset mesh.

(Refer to Annexure A for a summary report of the CFD simulation).

4.1.1 Mesh Configuration

The volume mesh in a simulation is a mathematical description of the geometry used in the problem being

solved. The mesh is constructed from mesh entities such as vertices, faces and cells [53].
45
The code used to construct the mesh was STAR-CCM+, as already specified. There are different meshing

techniques available to create these mesh constructions, but for the purpose of this study it was decided to

use the trimming method. The trimming method provides fast and high quality meshes [54]. This method

constructs hexahedral cells to represent the volume of the geometry.

Furthermore, the surface remesher and the prism layer mesher models were also activated. It is important

to note that the stretching mode for the prism layer mesher is set to wall thickness, while the rest of the

functions are kept to the default settings. In the case of the surface remesher, proximity refinement is

disabled as is automatic surface repair. The minimum surface quality must also be set to a value of zero

(0). Per region meshing is also set to true.

The cells on the xy-plane at z=0 was retained and the remaining volume cells in the 3-D mesh was removed

from the simulation. By doing this, the 3-D mesh was converted to a 2-D mesh (quadrilateral cells), thus

simplifying the simulation and reducing the CPU time required to run the simulation. To properly capture

the trailing edge flow, volumetric control regions were incorporated to refine the mesh behind the airfoil,

as shown in Figure 24. Volumetric control regions increases the number of cells and computation time

required substantially, but increases the accuracy of the solution [55].

Figure 24 – Quadrilateral 2-D mesh

Comparing the simulation results of an airfoil with the results in the literature, it is required to simulate the

airflow around the airfoil at a range of different angles of attack. An overset mesh was thus incorporated

into the simulation. This is illustrated in Figure 25, where the overset mesh is rotated through 8°. The
46
overset mesh function enabled the angle of attack to be changed by simply rotating the overset mesh region,

thus the airfoil, and running the simulation again.

Figure 25 - Overset mesh: airfoil rotated to an angle of attack of 8°

In order to capture transition, a very fine, layered mesh, called a prism layer, must be used on the boundary

wall of the airfoil (Figure 26).

Figure 26 - Prism layer on the leading edge of the NACA 0018 airfoil

The mesh can, however, not be too fine as this will mean more cells and therefore the simulation will take

longer to solve. A coarser mesh was therefore first generated and then refined.
47
4.1.2 Solver Configuration

In order to determine the aerodynamic coefficients of the airfoil, the solver had to be set up with flow

conditions matching those of the wind tunnel tests. Standard atmospheric conditions were accepted for the

simulation setup, as summarised in Table 3.

Table 3 - Atmospheric Conditions

Pressure 101325 Pa
Density 1.225 kg/m3
Temperature 288.16 K
-5
Dynamic Viscosity 1.85508x10 Pa.s

The wind tunnel tests from Timmer [52] were conducted at Reynolds numbers of 0.3x106 and 0.7x106. The

Reynolds number is calculated by

𝜌𝑉𝐷
𝑅𝑒 = 2. 19
𝜇

where ρ is the air density, V is the free stream velocity of the fluid, D is the length of the object and μ is the

dynamic viscosity of the fluid.

In the literature the pressure coefficient plot is also described for an angle of attack of 8° and a Reynolds

number 0.7x106. Therefore, the simulations were conducted at a range of angles of attack at a constant

Reynolds number value of 0.7x106. With a chord length of 0.25 m, the corresponding free stream velocity

should therefore be 40 m/s.

The boundary walls were kept stationary for the simulation and defined as non-slip. This was to simulate

the boundary walls of the wind tunnel.

To properly predict the aerodynamic coefficients, a transition model was used, as already noted. To

determine whether the solver had converged, the plots for these aerodynamic coefficients were also

monitored for convergence.

48
4.2 Validation Procedure

Before the simulation results can be compared with the experimental results, mesh refinement has to be

applied to determine whether mesh independence has been reached. Mesh independence refers to the point

where the solution is independent of mesh resolution, meaning that the solution is not improved any further

as the mesh is refined. As the mesh is refined, the number of volume cells is increased and therefore the

computing time required to solve the simulation. If the mesh is refined further without any significant

increase in the accuracy of the result, time is wasted. A base size should then be determined that will be

used to obtain the results which will be compared to the experimental results.

Mesh independence
0.013

0.012
Drag Coefficient [-]

0.011

0.01

0.009

0.008

0.007
0 100 200 300 400 500 600
Base size [mm]

Figure 27 - Mesh refinement for the 2-D simulation at an angle of attack of 0°

Referring to Figure 27, the drag coefficient decreases as the base size of the volume cells is decreased. This

occurs up until a certain point, at a base size of 150 mm, after which point the drag coefficient starts to

increase again. This is as a result of a too fine mesh. Reducing the base size, reduces all of the cells in the

model. Refinements should be incorporated in specific areas in order to properly capture the drag coefficient

of the model. The cells might become smaller than the tessellations of the model and this will cause the

49
solution to become unstable. It will have difficulty converging and therefore the drag coefficient value will

increase again.

Timmer [52] compared the experimental results from the wind tunnel with plots generated using RFOIL

[56]. RFOIL is a modified version of the MIT-developed XFOIL code and uses non-linear amplification

of the boundary layer [52].

The transition locations are given by Timmer and can therefore be compared with the results from the CFD

simulations. Transition has a large effect on the aerodynamic coefficients of the airfoil and will also

determine the accuracy of the simulation. Skin friction plots were used to determine the transition positions

in the CFD simulation and then compared with the literature.

In the literature it is also reported that at an angle of attack of 8°, laminar separation starts on the upper

surface. This is revealed in the pressure coefficient plot of the airfoil at an angle of attack of 8°, generated

using XFLR5. XFLR5 is a program with a user interface, based on XFOIL. As mentioned, RFOIL is also

based on the XFOIL code, thus it was assumed that the plots generated using XFLR5, would be nearly the

same as those generated using RFOIL, as in the literature. The plot of the pressure coefficient obtained

from the CFD simulation, was compared with reported findings in order to determine if laminar separation

occurred at the same point.

To further determine the accuracy of the CFD simulation, the values for the aerodynamic coefficients

obtained from the CFD simulation will be compared with the experimental graphs, as well as the graphs

obtained from RFOIL.

4.3 Validation Results

In order to ensure that the values used for plotting data was accurate, it was important to ensure that the

solution had converged (Figure 28). Monitor plots for the lift, as well as the drag coefficients, were used

to determine whether a converged solution had been reached.

50
Lift and Drag Coefficients
0.9

0.8

0.7

0.6
Coefficient Value [-]

0.5

Lift Coefficient
0.4
Drag Coefficient

0.3

0.2

0.1

0
0 500 1000 1500 2000 2500 3000
Iteration

Figure 28 - Lift and Drag coefficients, 8° angle of attack

In order to compare the CFD simulation results with the experimental and the XFLR5 results, the simulation

was run at angles of attack ranging from 1° to 16°. The simulations were also run using two different

turbulence models available in STAR-CCM+, namely the k-omega and k-epsilon turbulence models. The

airfoil overset mesh was rotated to change the angle of attack. Excel was used to plot the simulation data

against the experimental data.

51
Lift Coefficient against Drag Coefficient
1.6

1.4

1.2

Experimental data
CL

0.8
STAR-CCM+ k-omega
XFLR5
0.6
STAR-CCM+ k-epsilon

0.4

0.2

0
0 0.01 0.02 0.03 0.04 0.05 0.06
Cd

Figure 29 - Lift coefficient compared with the Drag coefficient, Re = 0.7x106

Referring to a study by Ahmed et al [55], it is indicated that the k-epsilon turbulence model is the model

typically used in simulations of external flow around bodies. It was therefore compared to the k-omega

turbulence model which was discussed in the literature survey.

When the lift coefficient is compared to the drag coefficient in Figure 29, both XFLR5 and STAR-CCM+

over-predicts the values for the lift coefficient, and the drag coefficient value increases with increasing

angles of attack. Even though this is the case, the same trend is obtained as in the case of the experimental

data. For lower angles of attack, the CFD data and the experimental data compare well. For drag, as well

as lift coefficient values for angles of attack up to 9°, the CFD and experimental values differ by less than

6%. At an angle of attack of 15°, however, the CFD data exceeds the experimental data in excess of 15%.

The k-epsilon turbulence model over-predicts the drag coefficient at low angles of attack; the value for the

drag coefficient being about 74% greater than the value calculated using the k-omega turbulence model.

52
The values for the lift coefficient were also compared with the different angles of attack. In Figure 30 it is

clear that the STAR-CCM+, as well as the XFLR5, plots compare extremely well with the experimental

data for low angles of attack.

Lift coefficient against angle of attack


1.6

1.4

1.2

1
CL [-]

Experimental
0.8
STAR-CCM+ k-omega
XFLR5
0.6
STAR-CCM+ k-epsilon

0.4

0.2

0
0 5 10 15 20 25 30 35
α [deg]

Figure 30 - Lift coefficient compared with the angle of attack, Re = 0.7x106

Again, at angles of attack higher than 10°, STAR-CCM+ and XFLR5 over-predict the values for the lift

coefficient, exceeding the experimental data by more than 15%. This is the case for both turbulence models

used in the STAR-CCM+ simulations. The same profile is still maintained. However, the values for STAR-

CCM+ never differ by more than 5% from the XFLR5 data.

The experimental tests were completed to an angle of attack of 30°, but the values for XFLR5 only extended

to 24°, after which the result no longer converged. At an angle of attack higher than 16°, STAR-CCM+

had difficulty converging, oscillating around a single value. This is because the flow around the airfoil has

53
separated from the surface. This separation point is indicated in Figure 31. This eliminates the need for

the prediction of transition as it has already occurred and the solution will therefore not converge.

Separation point

Figure 31 - Separation point on airfoil

The STAR-CCM+ values are therefore only plotted until an angle of attack of 16°. The lift coefficient at

the different angles of attack is accurately calculated by both turbulence models used in STAR-CCM+, with

the values remaining within 2% of one another.

It is indicated by Timmer that at 0° angle of attack, transition occurs at 54% of the cord length (c) of the

NACA 0018 airfoil [52]. In Figure 32 transition occurs where the skin friction coefficient starts increasing,

before decreasing later on. Here it occurs at about 0.14 m, which is 56% of the chord length. This is 2%

less than the experimental value.

54
Skin Friction Coefficient Plot
30

25

20
CSF [-]

15

10
Start of transition

0
0 0.05 0.1 0.15 0.2 0.25 0.3
x [m]

Figure 32 - Skin friction coefficient plot for a 2-D NACA 0018 airfoil, 0° angle of attack

In Figure 33 it is observed that the pressure coefficient values for STAR-CCM+ compare very well with

the data obtained from XFLR5. The plot for XFLR5 data runs on the line of the STAR-CCM+ data plot,

remaining within 5% of one another.

55
Pressure Coefficient Plot
Position [-]
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
-2.6
-2.4
-2.2
-2 Transitional
-1.8
-1.6 separation bubbles
-1.4
-1.2
XFLR5
-1
CP [-]

-0.8 STAR-CCM+
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
1.2

Figure 33 - Pressure coefficient of the 2-D NACA 0018 airfoil, 8° angle of attack

Start of separation

bubble in either

case

Figure 34 - Separation bubble formation on the upper airfoil surface, 8° angle of attack

In Figure 34 the separation bubble is indicated by the slight rise in the pressure coefficient for XFLR5.

Here the separation bubble formation starts to form at about 15% of the chord length. Looking at the same
56
figure, the data from STAR-CCM+ reveals that the separation bubble starts a bit earlier, at about 13% chord

length. Note that in Figure 33 and Figure 34, the length of the airfoil is not expressed in the actual chord

length of 0.25 m, but as a percentage.

4.4 Conclusion

The CFD results compare very well with the experimental results from the literature at low angles of attack.

The difference observed is that the CFD results are over-predicted at higher angles of attack. After

separation has occurred, the CFD software has difficulty converging and the results after separation can

therefore not be compared to the experimental results. The transition locations of both sets of results also

compare well, keeping the above in mind. It can be concluded that the use of the specific transition models

for simulating transition of 2-D profiles yields satisfactory results. The transition models can therefore be

validated for a 3-D case similar to that of the solar-powered vehicles.

57
Chapter 5: 3-D Validation: HPT VeloX1

In order to improve on the current SV design of the North-West University, it is necessary to analyse the

current design using CFD. To validate the procedure that would be used in order to analyse the current

design, CFD is used to analyse the HPT’s (Human Power Team) VeloX1 bicycle (see Figure 35 [57]). The

HPT is a group of students from the Delft University of Technology in the Netherlands, who aims to use

this bicycle to break the land speed record for this class of human powered vehicles.

Figure 35 – Human Power Team’s VeloX1 [57]

A full scale model of the VeloX1 had been analysed in a wind tunnel at a velocity of 28m/s, and these

results were made available by the HPT group to use for validation purposes of this study. The road test

data was also supplied. It was therefore viable to do a CFD analysis on the CAD model of the VeloX1 and

compare it with the experimental data.

Once the simulation is validated, the same procedure would be applied to the North-West University’s SV

to determine the aerodynamic drag values of the vehicle. This could be done as the driving speeds of solar-

powered vehicles could also be of the order of 28 m/s, which is about 100 km/h.

58
The experimental results for the road tests as well as for the wind tunnels tests, were obtained from the team

and permission was granted by the team to use the model in the validation process for this study.

5.1 Validation Setup

In order to properly validate the 3-D model, transition locations have to be accurately determined for 3-D

flow around the HPT VeloX1 model. The transition locations of the CFD simulations will be compared

with the wind tunnel transition locations, thus ensuring that the CFD simulation is accurate enough to use

as a design tool for the design of the SV. Furthermore, the drag coefficient values of the simulation should

also be relatively close to the wind tunnel results to further enhance the validity to use the CFD simulation

as a design tool.

Once again, as in the case of the 2-D validation, STAR-CCM+ will be used to do the CFD simulation.

During the 2-D validation, the γ-Reθ transition model, coupled with the k-ω SST turbulence model, was

used to determine the transition on the airfoil. These positions position were determined within 5% of the

experimental values, and will therefore also be used to determine the transition locations on the 3-D model.

Regarding the geometry to simulate the fluid flow around the bicycle in the wind tunnel, a rectangular

tunnel was used based on dimensions described in the literature, originally by Mansor et al. [27]. The

dimensions of the simulated wind tunnel was not the same as the experimental wind tunnel used. A rough

simulation was conducted with dimensions similar to the actual wind tunnel, but this increased the number

of cells required in the simulation without a significant improvement in the results obtained. It was thus

opted to use the simpler, rectangular wind tunnel in the simulation. The rectangular tunnel should be at

least twice as high as the model, with the outlet boundary at least five vehicle lengths behind the model to

ensure that the wake of the vehicle is properly captured. The inlet boundary should be at least two vehicle

lengths ahead of the model to ensure that the flow has had sufficient time to re-laminarise before reaching

the model.

59
Symmetry Plane

Velocity Inlet

Pressure Outlet

VeloX1 Model

Figure 36 - Domain description for the 3-D validation case

The boundary conditions for the 3-D validation case is described in Figure 36. The inlet boundary was set

to a velocity inlet as in the 2-D validation setup, with a velocity of 28 m/s. The pressure outlet is set to a 0

Pa pressure outlet condition and the walls of the tunnel was treated as fixed walls to represent a wind tunnel.

Due to the symmetric nature of the model, only half of the geometry was simulated. This meant that the

results would have to be doubled to get the value for the whole model. The plane separating the domain in

half is treated as a symmetry plane.

5.1.1 Mesh Configuration

Different meshes were compared in order to determine which was most accurate. It is generally accepted

that the results are more accurate when using the polyhedral mesher instead of the trimmer mesher. The

trimmer mesh uses hexahedral cells to build the geometry, while the polyhedral mesh uses polygons with

up to 14 sides. The polyhedral mesher, however, uses more cells to build the geometry. The trimmer mesh

uses much less than the other meshing techniques and therefore less computing time. The trimmer mesh is

therefore used to conduct the simulations.

60
Figure 37 - VeloX1 geometry

As for the 2-D validation case, the surface remesher and the prism layer mesher models were also activated.

The stretching mode for the prism layer mesher is set to a stretch factor while the rest of the functions are

kept to the default settings. In the case of the surface remesher, proximity refinement is disabled, as is

automatic surface repair. The minimum surface quality must also be set to a value of zero (0).

To properly capture the trailing edge flow, volumetric controls were once again incorporated in the mesh

to refine the mesh behind VeloX1. This is indicated by the higher density of cubes, as shown in Figure 38.

Figure 38 - Mesh refinement for VeloX1 trailing edge flow

In order to capture transition, a prism layer must again be used on the boundary wall of the VeloX1, as in

the 2-D validation case with the airfoil.

61
Figure 39 - VeloX1 prism layer on surface mesh

5.1.2 3-D Solver Configuration

Atmospheric conditions given in Table 3, were also used in the 3-D validation.

The wind tunnel tests were conducted at velocities of 28 m/s and therefore the inlet boundary velocity was

set to the same value. This resulted in a Reynolds number of almost 5x106. This Reynolds number is based

on the length of the bicycle, which is 2.7 m.

5.2 Validation Procedure

To determine whether the 3-D validation is accurate, the transition locations of the simulation will be

compared with the transition locations detailed in the literature, as supplied by the HPT. This literature is

in the form of a report and gives the transition locations for the wind tunnel tests, as well as a simulation

that was conducted using VSAero.

VSAero is based on the assumption that the model being simulated is at rest in a moving airstream, relating

to the chosen method to conduct the simulations of the model, which is also simulated in a moving airstream

rather than moving at a velocity through the air. The flow field is also assumed to be inviscid, irrotational

and incompressible except for regions that are dominated by viscous, rotational flow. Such areas are

confined to thin surface boundary layers and wakes. After reviewing the program, Maskew concludes that

VSAero produced comparible results to high order methods [58]. These higher order methods include

modern day CFD methods.


62
The HPT also gave the drag area values for the wind tunnel tests, the road tests and the drag area (CDA) as

calculated by VSAero. This means that the tests were not conducted using a reference area. The CFD

simulations will therefore be run to give only the resulting aerodynamic drag force (D) on the surface of

the model:

1
𝐷 = ( 𝜌𝑉 2 ) (𝐶𝐷 𝐴) 5. 1
2

This will give a value for the drag area, which will then be compared with the experimental values.

The pressure plots on the surface of the vehicle body will also be compared with the experimental results.

These plots will also give an indication of the accuracy of the simulation.

5.3 Validation Results

To determine whether the CFD simulation of the VeloX1 is accurate, the simulation results must be

compared with the experimental data as described in the preceding section 5.2.

The data supplied by the HPT describes the VeloX1 as having a drag area value, for the wind tunnel test,

of 0.031 m2 and a value of 0.033 m2 for the road tests. These tests were conducted without closing the

wheel holes and the ventilation ducts. The wind tunnel tests were repeated with the ducts closed off. This

yielded a CDA value of 0.025 m2. The CAD model does not make provision for these ducts or wheel

openings, therefore the CFD simulation results must be compared with the second case (when the ducts

were closed).

The VeloX1 has a frontal area of 0.363 m2. Using this value, it is easy to determine the drag coefficient of

the VeloX1 when the ducts are closed off. With the drag area of 0.025 m2, the experimental drag coefficient

is 0.069.

To determine if the CFD simulation has converged, a force monitor plot is generated and displayed in

Figure 40. This is the aerodynamic force (D) that is exerted on the surface of the VeloX1 and Eq. 5. 1 can

be used to determine the simulated value for CDA. The plot shows that the results oscillate around a final

value. This can be eliminated by averaging the values of the last 100 iterations, which will give a fair

representation of the mean drag force value.

63
Aerodynamic Force on VeloX1
40

35

30

25
Force [N]

20

15

10

0
0 250 500 750 1000 1250 1500 1750 2000
Iteration [-]

Figure 40 - STAR-CCM+ force monitor plot

The simulated value for half of the aerodynamic force on the VeloX1 is 4.648 N. Because a half

symmetrical body was simulated, this value was multiplied by 2 resulting in 9.296 N for the whole body,

which translates to a CDA value of 0.0196 m2 and a drag coefficient of 0.0533. This is 22.5% less than the

experimental value from the wind tunnel tests. It is difficult to simulate the aerodynamic drag precisely,

but predicting the transition location is much more important. The difference can be attributed to features

in the wind tunnel that are near impossible to incorporate into the CFD simulation. CFD simulations were

also run where effort was made to draw the wind tunnel geometrically correct, but this yielded a difference

of less than 5% to the rectangular wind tunnel geometry. The difference noted here was that the rectangular

wind tunnel had fewer cells. With this small difference in drag force values, it was opted to use the

rectangular geometry due to its simplicity and shorter computing time.

64
Table 4 - Drag Area Results Comparison

Model CDA CD

Wind Tunnel Test 0.025 0.0688

VSAero 0.015 0.0413

STAR-CCM+ 0.0196 0.0533

Looking at the simulations conducted using VSAero, the CDA value is 0.015 m2 and the drag coefficient is

0.0413. The geometric setup for the wind tunnel of this simulation is not known, but the reasons for this

lower prediction will probably be the same as for the above-mentioned. This value is about 22.5% lower

than the STAR-CCM+ simulated value.

The transition locations were determined by the HPT using wind tunnel tests, as well as using VSAero. .

Laminar flow is experienced on the front half of the VeloX1’s body. At about the halfway mark, the laminar

boundary layer starts to undergo transition to a turbulent boundary layer. Referring to Figure 41, the point

of separation is indicated by an arrow. Transition is induced at the start of the front wheel fairing, as well

as at the top of the windscreen. The position of the ventilation duct also causes the flow to separate

prematurely.

Transition line

Figure 41 - VeloX1 transition location with the ventilation ducts taped [57]

The transition location of the model simulated using VSAero, shows quite a few differences to the wind

tunnel results. The model in Figure 42 shows that the flow does not separate at the start of the front fairing,

neither does the location of the windscreen as the whole model has a smooth surface.

65
Transition line

Figure 42 – VeloX1 transition location by VSAero [57]

Referring to Figure 43, transition determined by CFD occurs just before half the length of the model, with

laminar boundary layers on the top of the model progressing past half the length of the model. One can see

that the CFD-predicted separation is a combination of the wind tunnel results and that of VSAero. The

geometry used does not have a windscreen to cause flow separation, observed when referring to Figure 41.

However, the laminar flow extends past the halfway mark, indicated by the solid line in both Figure 42 and

Figure 43.

Transition/separation line

Figure 43 - VeloX1 transition location by STAR-CCM+, turbulent kinetic energy plot

The pressure coefficient on the surface of the VeloX1 was also supplied by the HPT and will be compared

with the simulation results. The pressure coefficient calculated by VSAero is illustrated in Figure 44.

66
Figure 44 - HPT VeloX1 pressure coefficient by VSAero

Comparing Figure 44 with Figure 45, it can be observed that the pressure on the surface of the VeloX1 is

very similar, with differences in the maximum and minimum values. However, the minimum and

maximum pressure locations are the same for both simulations, with the lower pressure values being around

the centre of the vehicle body, thus near the thickest section. This correlates with the literature, which

indicates that flow separation will usually occur when the pressure reaches a minimum and pressure

recovery occurs to the rear of the vehicle body.

Figure 45 - VeloX1 pressure coefficient by STAR-CCM+


67
It is also clear, when referring to Figure 45, that total pressure recovery is not achieved, indicating that flow

separation has occurred.

5.4 Conclusion

The values calculated by the CFD simulation for the 3-D case are lower than that of the experimental results,

22.5% lower in the case of the STAR-CCM+ simulation and about 50% lower for the VSAero simulations.

This can be attributed to various reasons, including the fact that the actual model has a windscreen, wheel

holes and ventilation ducts. All of these factors contribute to a higher drag value. Other factors to consider

include surface roughness and manufacturing irregularities. The transition plots of the different models

compare relatively well. There are differences in the extent to which the models experience laminar flow,

but these differences can be attributed to differences between the actual model and the CAD models (the

CAD model being smoother). The pressure distribution plots also compare well. It can therefore be

assumed that the transition models can be used to predict the transition of the 3-D models.

68
Chapter 6: Conceptual Design

During the conceptual design phase, different ideas for the design were explored and references were made

to existing designs of popular solar vehicles (Refer to Section 2.6). These steps are important during this

phase, leading up to the final design of the vehicle. Many existing designs have had the driver pod in the

middle of the vehicle, with the wheels covered by fairings. However, by analysing these designs in terms

of the design parameters discussed, it is clear that the frontal area of the design is significantly increased.

It was suggested to move the driver to the side and integrating the driver pod with the fairing, resulting in

a catamaran design. This not only resulted in a design with a smaller frontal area, but also one with a

smaller wetted surface. Theoretically this means a lower overall drag coefficient. This also decreases the

junctions between the fairings and the body and should further aid in decreasing the drag coefficient of the

vehicle, which is a design parameter. These parameters can be further reduced by using shorter and

narrower fairings on the wheels opposite to the driver position of the vehicle.

Three basic designs were proposed for the next generation solar-powered vehicle based on the above ideas.

These designs were proposed to the team for comparison and discussion in order to determine which design

will potentially have the best efficiency. Review of existing solar-powered vehicles revealed that the

proposed concept designs were similar to these existing designs. This indicates that the methodology used

to achieve these concept designs is also used by the leading solar car teams around the world.

6.1 Concepts

The airfoil used for the main body of the vehicle is a modification to the S5 airfoil used in the wing design

of Jonker Sailplanes. This airfoil was obtained from Jonker Sailplanes for use in the design of the solar-

powered vehicle and has also been used in the 2012 vehicle. This also applies to the airfoil used for the

wheel fairings during the conceptual phase, which is the symmetrical airfoil used for the tail wing of the

same sailplanes. By using the same airfoil for all three conceptual designs, the design with the least amount

of aerodynamic drag can be chosen, knowing that the airfoil will have the same properties.

69
6.1.1 Concept based on Symmetry

The first concept proposed would be similar to the 2012 solar powered vehicle, with the driver pod in the

middle of the vehicle. Opposed to the previous vehicle design, the driver would be seated to the front of

the vehicle, thereby aiding in the aerodynamic stability of the vehicle. Furthermore, each wheel would be

covered individually by a fairing with the suspension, steering and braking systems enclosed within the

fairings. This reduces the frontal area of the vehicle slightly, compared to the 2012 solar powered vehicle.

Refer to Figure 46 – First concept design (symmetrical).

Figure 46 – First concept design (symmetrical)

As the driver is seated in the centre of the vehicle, this design results in a larger frontal area, as well as a

larger wetted surface. In theory this already indicates that this design will have a larger drag coefficient

when referring to Eq. 2.2. This equation states that when the velocity of the air flowing past the object or

vehicle is kept constant, the frontal area and the drag force will be directly proportional. If the frontal area

is increased, then the drag force will also increase proportionally.

6.1.2 Asymmetrical Concept with Long Fairings

It was discussed in the previous sub-section that having the driver seated in the centre of the vehicle will

have a negative impact on the design because the frontal area of the vehicle will be too large and therefore

70
the drag of the vehicle will also be substantially higher. The second concept design is therefore modified,

moving the driver to the right hand side of the vehicle, between the front and rear wheel. This means that

the driver pod and the right hand wheel fairings are integrated into a single structure, thereby reducing the

frontal area of the vehicle substantially. Once again referring to Eq. 2.2, the drag force of the vehicle will

be reduced as the frontal area of the vehicle is reduced.

Figure 47 – Two long fairings with driver to the right hand side of the vehicle

This is the catamaran effect mentioned during the brainstorming process. The wheels on the left hand side

of the vehicle would also be covered by a single, long fairing (as depicted in Figure 47), but one that is

slightly narrower than the right hand side fairing (Figure 48). This basic design is also used by Nuon Solar

Team, reviewed during the literature survey.

71
Figure 48 – Front view of the second concept

This would allow components to be installed inside the fairing in a compartment between the wheels on the

left hand side. By using this space between the wheels, the main body can be narrower and thus the frontal

area can be reduced even further. This means it may have a clear advantage over the first concept as the

frontal area would be significantly reduced. It should, however, be noted that by utilising the space between

the wheels on the left hand side of the vehicle, the centre of gravity of the vehicle will shift to the rear, even

though the driver is still seated forward of the neutral point of the vehicle. This might have implications

for the aerodynamic stability of the vehicle. Most of the weight should be towards the nose section of the

vehicle, moving the centre of gravity as far forward from the neutral point as possible.

6.1.3 Asymmetric Design with Short Fairings

The third concept idea is based on the same principal as the preceding concept, with the driver seated in a

single long fairing on the right hand side of the vehicle. The difference between the two concepts is that

the third concept sees a change in the left hand fairing design. The single long fairing is replaced by two

shorter fairings as in the symmetric design proposed first (Figure 49).

72
Figure 49 – Asymmetric design with three fairings

This change means that the components will have to be installed inside the main body of the vehicle as

there is not enough space in the shorter fairings. It will then be possible to install the components to the

front section of the vehicle and thereby moving the centre of gravity further forward from the neutral point

of the vehicle. This will increase the aerodynamic stability of the vehicle, meaning it will be less likely to

experience lift on the nose of the vehicle when driving. It also means that the front wheels will experience

better traction and therefore the vehicle will have better road handling capabilities at higher speeds.

Figure 50 – Front view of the third design, showing the effect that the shorter fairings has on the frontal area

73
By using shorter fairings, the maximum width of these fairings will be much narrower (Refer to Figure 50)

than when a single longer fairing is used. This means that the frontal area, as well as the wetted surface of

the third design, will be even less than in the second concept, which should result in a further reduction of

the aerodynamic drag force.

Theoretically it therefore means that the third design should be the most aerodynamic and it will therefore

make sense to use this design for optimisation. This design is also similar to the design used by Tokai

University.

6.2 Comparison Setup

During the validation process described in the preceding chapter, Chapter 5: 3-D Validation: HPT VeloX1,

a setup and procedure was described which should provide a suitable platform on which to conduct the

comparison of these concept designs. Comparing the concepts is an essential step in the design process of

the vehicle. By doing an aerodynamic study about the drag force of the concept designs, it is possible to

compare the different designs and choose the design with the most favourable aerodynamic properties. It

should be noted that for the purpose of this dissertation, to accurately determine the aerodynamic drag of

the vehicle, the transition location should be determined. The turbulent kinetic plots will therefore be used

to show transition locations and compare it with one another. The drag area of the models, along with the

drag coefficients, are also good indications of which concept design can be used in the detail design phase.

The chosen concept will be refined to further reduce the aerodynamic drag of the vehicle.

6.3 Comparison Procedure

To attain a value for the aerodynamic drag coefficient, a force monitor was used in the STAR CCM+

simulation to measure the drag force of the vehicle body. To speed up the comparison process, a server

with 64 processors were used. Equation 5. 1 is again used, as in the 3-D validation case in Section 5.2.

This will give a value for the drag force (D), so that the drag area (CDA) can be determined. The frontal

area of the vehicle body (A) is also known and this will provide the value for the aerodynamic drag

coefficient (CD). The same settings were used as described in sections 5.1.1 and 5.1.2 of the 3-D validation

procedure.
74
6.4 Comparison Results

Comparing the different concepts, following the process as described, yielded the results summarised in

Table 5, where C1, C2 and C3 represents the three concepts discussed (respectively).

Table 5 - Conceptual design comparison

C1 C2 C3
F 57.9 46.8 43.9
CdA 0.121 0.098 0.091
A 0.864 0.812 0.766
Cd 0.140 0.120 0.119

Referring to Table 5, it is quite clear that the least aerodynamic design would be the first concept (C1) being

a symmetric design. C1 has a high drag coefficient of 0.14, largely due to the larger frontal area as well as

a larger wetted surface. This means that there is more air in contact with the surface of the vehicle, thus a

higher shear force. The increased number of appendages increases the interference drag between the body

parts. The higher the shear force and the interference drag on the surface of the vehicle, the higher the drag

force will be and essentially the drag coefficient.

Figure 51 - Turbulent kinetic energy of the first concept


75
Referring to Figure 51, it is observed that C1 experiences substantial laminar flow on the front fairings.

There is still laminar flow on the rear fairings, but increased turbulent flow is observed. When the vehicle

is viewed from a lower angle (Figure 52), it is clear why the drag force of the vehicle is so much higher in

than the other concepts.

Flow transition

Horseshoe vortices formed

Large wake

Figure 52 - Bottom view of the symmetric concept vehicle

The driver pod increases the wetted surface substantially and produces a very large turbulent wake due to

large vortices that formed at the pod junction. The rear fairings also experience increased turbulent flow

on the inner sides next to the driver pod, as this narrower section increases the velocity of the air and thereby

increases turbulence. The pod also shows quick transition of laminar flow.

The second concept design (C2) shows a substantial reduction of almost 20% in the drag force. This is

because of the reduction in the frontal area of the vehicle. Other factors contributing to this large reduction

is the smaller wetted surface as well as fewer appendages, reducing the number of junctions. Fewer

junctions mean less interference drag. The drag coefficient of this model is 0.12, which is about 16% less

than the first concept and this is because of a reduction in the frontal area of about 6%.

76
Smaller turbulent wake

Smaller vortices

at start of fairings

Figure 53 - Turbulent kinetic energy of the belly section of the second concept

In Figure 53 it can be observed that the longer fairing experiences laminar flow to about two thirds of the

fairing length on the outside and to about half of the fairing length on the inner side, helping to reduce the

drag of the vehicle further. The effect of the horseshoe vortices formed by the fairings are much smaller

than in the previous concept. This means smaller wakes will be created behind the fairings, all contributing

to the lower drag force experienced with this concept.

As discussed, the third concept design (C3) is expected to be the most aerodynamic as the wetted surface

and the frontal area is further reduced from C2. C3 has a frontal area which is about 5.7% smaller than C2

but this relates to only a 0.7% reduction in the drag coefficient to a value of 0.119.

77
Higher levels of turbulent flow

Larger wake

Figure 54 - Turbulent kinetic energy of the belly section of the third concept

This small reduction is attributed to the high levels of turbulent energy experienced on the rear fairing and

also a larger wake behind this same fairing. This means that, even though the frontal area of the vehicle is

reduced, the drag area of the vehicle is larger and this means larger areas of turbulent flow. The vortices

created at the start of the fairings, do not differ much from C2.

6.5 Conclusion

Looking at the results, it can be argued that the small difference between C2 and C3 means that either

concept can be modified and improved on in the detail design phase. This is true but because C3 has a

much lower frontal area, it means that if the interference drag and turbulence on the rear fairing is improved,

the drag can potentially be reduced more than when the same procedures are applied to C2. This increases

the advantages of the third concept design. Keeping in mind the descriptions of the concepts regarding the

placement of internal components, it was argued that C3 holds even greater advantages in that the

aerodynamic stability potential of this concept is the greatest. This, coupled with the fact that it did have

the lowest drag of all the concepts tested, means that the third concept holds the greatest potential for

improvement and was selected for the detail design phase.

78
Chapter 7: Detail Design

After the baseline had been determined, the different features of the design which were discussed, had to

be optimised in order to reduce the drag force of the vehicle. These features included variations in the

airfoil profiles used for the fairings, the canopy and the body of the vehicle. Individual simulations of the

different components could determine the more aerodynamic shapes to be integrated with the final vehicle

body. This is done while constantly considering the competition regulations.

7.1 Component Analysis

7.1.1 Analysis Procedure

In order to compare the different components with each other, either 2-D simulations or 3-D simulations

were used. As the design of the different components was based on airfoil profiles, the first step in the

comparison was to evaluate different airfoil profiles in 2-D using XFLR5. The airfoil profiles were chosen

such that the transition location was as far to the rear as possible. This theoretically implies that the laminar

flow on the 3-D profile will be extended further.

For the 3-D simulation the same simulation setup as discussed in Chapter 4 was applied to the components

in STAR-CCM+. The mesh and solver configurations were therefore already validated and it was assumed

that the transition locations on the individual components would give a fair representation of the laminar

flow for the vehicle. The components were compared with each other by evaluating the drag coefficients,

as well as studying the transition locations on the turbulent kinetic energy plots. On these plots, the

transition location is where the colour in the figure abruptly changes, from a dark shade (dark blue) to a

lighter shade (yellow or light blue).

The results of these comparisons are described in the following sub-sections.

7.1.2 Main Vehicle Body

The design of the main body of the vehicle was based on a modified version of the JS1 sailplane wing

airfoil (Figure 55). This is testament that it was an excellent profile to be used to enhance the aerodynamic

properties of the vehicle.

79
Modified JS1 Sailplane Wing Airfoil
0.1

0.05

-0.05

-0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 55 – Modified JS1 sailplane wing profile used for main body

Main Body CP Curve


Position [m]
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
-0.15

-0.1

-0.05

0
CP [-]

0.05

0.1 Transitional

separation bubbles
0.15 XFLR5
STAR-CCM+

0.2

Figure 56 - Pressure coefficient curves of main body profile

Referring to the above figure (Figure 56), the XFLR5 3-D results at a 0° angle of attack shows that the body

experiences transitional separation bubbles at about 35% of the chord length on the upper surface. The

laminar flow extends a bit further on the lower surface, to about 50% chord length. These separation

bubbles cause the laminar flow to transition. The transition locations in the STAR-CCM+ simulation is

calculated to be about 35% chord for the upper and lower surfaces. This compares well with the XFLR5

results.

80
Referring to the STAR-CCM+ 3-D pressure coefficient plot of the main body in Figure 56, the same

pressure gradient is perceived, with slightly higher pressure coefficients. The minimum pressure coefficient

values in the two cases differ by about 20%. The minimum value in the case of the STAR-CCM+ results

is about -0.08 compared with almost -0.1 in the XFLR5 results.

The peak in the pressure coefficient value at the trailing edge of the 3-D case, is a result of the trailing edge

not having a knife edge, but a thickness of 5 mm. The solution, therefore, does not completely converge

and the spikes in the above figure is observed. Regarding the stability of the airfoil, the pitching moment

coefficient (Cm) is -0.006, calculated by XFLR5. This differs greatly from the value calculated in STAR-

CCM+, which has a value of -0.018 for Cm. This difference might be attributed to calculation methods used

by the two codes, but the more probable reason might be the geometric differences. The body used in

STAR-CCM+ has rounded sides, while the body used in XFLR5 has flat sides. The vortices created on the

body sides might alter the moment forces on the body differently. As discussed in the literature, the pitching

moment must be negative in order to counteract the effect of the lift force. This means that the airfoil will

be statically stable.

This airfoil was made into a 3-D part (Figure 57), with large rounded edges to improve the ease of

manufacture, as well as providing a large flat surface where internal components of the vehicle, such as

batteries, could be installed.

Figure 57 - Main body of the vehicle

Note the rounded leading-edge of the main body in Figure 58. This has a large radius which not only

improves the manufacture of the moulds and the actual parts, but helps to improve the aerodynamics by

81
avoiding sharp edges, as was the case of the 2012 vehicle. If sharp edges were present on the leading edge,

it would cause the flow to separate from the surface. The rounded surface has the effect of easing the flow

over the body of the vehicle. This premature separation would result in a higher drag force and therefore

negatively affect the aerodynamics. This means that the laminar flow would extend past the leading-edge

(Figure 59). In the case of sharp corners on the leading-edge of the body, flow would have transitioned and

minimal laminar flow would have been possible on the sides of the main body.

Figure 58 - Rounded leading-edge of the main body

Transition location

Figure 59 - Laminar flow on leading-edge of main body

7.1.3 Fairing Profile Selection

It is easier to select an airfoil from a selection of profiles that is known to have laminar flow and has already

been proven in a 3-D case. In this case the airfoil was a modified version of the tail wing of the JS1
82
sailplane, namely the RUD20 airfoil (Figure 60). This sailplane has been proven to be the best during

multiple world championships, and the laminar flow properties are favourable for the application of wheel

fairings.

RUD20 Airfoil
0.1

0.05

-0.05

-0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 60 - RUD20 airfoil

For the RUD20 airfoil, the point of highest pressure is near 50% chord length, which is favourable. This

indicates that the laminar flow will also extend along the length of the profile, with the transition location

at approximately 45.8% of the chord length. The RUD20 airfoil has a drag coefficient value (Cd) of 0.005.

This airfoil was compared with other aerofoils exhibiting the same characteristics. These aerofoils had to

be symmetric, so that the wheels of the vehicle would be centred in the fairing. The trailing section could

not draw too sharply into a narrow trailing edge, because this would limit the space available in the fairing

for the wheels, suspension, etc.

The YS900 airfoil, shown in Figure 61, was compared with the RUD20 airfoil because its maximum

thickness was further back at 51.91%. YS900 also had a stronger pressure recovery which would make it

better suited when two fairings were in tandem and the rear fairing was in the slipstream of the front fairing.

YS900 Airfoil
0.1
0.05
0
-0.05
-0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 61 - YS900 airfoil


83
The transition location of the YS900 airfoil is much further back than the RUD20 airfoil, at 88.5% of the

chord length. This is extremely favourable and it also has a Cd value of only 0.003. These characteristics

made the YS900 airfoil better suited to the fairings design.

3-D CAD parts were made using these profiles, taking into account the points discussed in section 2.4,

namely the inclination angle of less than 10°. These 3-D parts were integrated with the rough 3-D concept

model of the vehicle selected in the previous chapter and the model of the vehicle was re-evaluated in

STAR-CCM+ and compared to each other. The RUD20 fairing had a drag force value of 73.32 N. The

YS900 fairing, as expected, had a lower drag force value of 72.3 N due to the extended laminar flow, albeit

only marginally.

Another airfoil with a favourable profile is the NACA 16015 airfoil (Figure 62). This airfoil is closer to

the RUD20 airfoil although the trailing section does not draw into thin point which is better for the fairing

design.

NACA 16015 Airfoil


0.1

0.05

-0.05

-0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 62 - NACA 16015 airfoil

The thickest section of the wheel is at 50% chord length. As the wheel will usually be centred in the fairing,

it makes sense from a manufacturing standpoint that the fairing should be thickest at 50% chord length.

However, the transition characteristic of this airfoil was not favourable.

To thicken the airfoil at the rear for reasons stated, the NACA 16015 was integrated with the YS900 airfoil

and shown in Figure 63. This resulted in an airfoil with a Cd value of 0.003 and a transition location of

about 75% of the chord length. The pressure recovery was not as drastic as that of the YS900 airfoil, but

stronger than the NACA 16015.

84
50:50 NACA 16015 and YS900 Combination Airfoil
0.1

0.05

-0.05

-0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Figure 63 - NACA 16015 and YS900 airfoils combination

NACA 16015 and YS900 Combination Airfoil CP Curve


Position [-]
0 0.2 0.4 0.6 0.8 1
-0.4

-0.2

0
CP [-]

0.2

0.4

0.6

0.8

Figure 64 - Pressure distribution along the chord length of the NACA 16015 YS900 airfoil

The NACA 16015 YS900 airfoil exhibited poorer transition in the 2-D case, but when the 3-D CAD fairing

was made and evaluated using CFD, the overall drag force value for the fairing was 71 N. This was better

than any of the other fairings evaluated due to the favourable pressure gradient up to about 70% chord

length (Figure 64). The drag force value for this fairing was reduced even further by reducing the width of

the fairings to the minimum width, thereby reducing the frontal area, which was one of the design

parameters. With this modification the drag force value of the vehicle was 64 N.

85
7.1.4 Fairing Design

As mentioned in the previous section, the rear wheel fairing is in the slipstream of the front fairing. The

front fairing would ideally have a constant pressure gain with a flatter profile. Refer to Figure 64. The rear

fairing should have a much steeper gradient with a very sharp pressure recovery. Theoretically this should

enhance the laminar flow on the rear fairing.

3-D CAD configurations of the fairings were made and evaluated using CFD to determine if this was the

case.

Configuration 1

The first configuration had the front fairing as the NACA 16015-YS900 combination fairing. However,

due to the YS900 fairing having a steeper pressure recovery, it was selected as the rear fairing. This

configuration is depicted in Figure 65.

Figure 65 – Fairing Configuration 1

86
Re-attachment of flow

CP section plane

Figure 66 - Transition locations of flow on fairing Configuration 1

This configuration had a drag force of 18.24 N. Even with the sharper leading-edge and sharp pressure

recovery of the YS900 airfoil in the rear, the flow transitioned relatively quickly. Figure 66 shows the

turbulent kinetic energy plots and the transition locations on the fairings. The reason for the poor laminar

flow on the rear fairing was because the fairing is already in a turbulent flow region.

Configuration 1 CP Curve
Position [m]
0 0.5 1 1.5 2 2.5 3 3.5 4
-0.6

-0.4

-0.2

0
Front Fairing
0.2
CP [-]

Rear Fairing

0.4

0.6

0.8

1.2

Figure 67- Pressure coefficient plot of Configuration 1

87
It is clear that, referring to the turbulent kinetic energy plots in Figure 66, the flow re-attaches in the pressure

recovery region of the rear fairing. This is indicated by the darker coloured sections on the trailing edge.

Even though this is the case, looking at the pressure coefficient plots in Figure 67, the pressure recovered

is not all that satisfactory. Therefore, a larger wake will be present behind the rear fairing. The section

plane from which the pressure coefficient data is extracted, is indicated by the horizontal line in the above

figure.

Configuration 2

To test if a much sharper leading-edge and an even stronger pressure recovery for the rear fairing would

further enhance the laminar flow on the rear fairing, the following configuration was made, while keeping

the front fairing the same as in Configuration 1. The front fairing was kept the same because it showed

good laminar flow in configuration 1, so there was no reason to substitute it at this time.

Figure 68 – Fairing Configuration 2

88
Figure 69 - Transition locations of flow on fairing Configuration 2

This configuration (Figure 68), however, had a higher drag value at 19.44 N due to the higher levels of

turbulent kinetic energy on the rear fairing (visible on the turbulent kinetic plot in Figure 69 as red). This

was an indication that the sharper leading-edge did not promote much laminar flow, possibly again due to

the already turbulent flow. The rear fairing was also quite close to the front fairing, meaning that the flow

did not have enough time to re-laminarise before reaching the rear fairing. This configuration is not

favourable; not only due to the higher drag value, but also because of the limited space in the rear fairing

for components such as the suspension and the braking mechanism on the rear wheel.

89
Configuration 2 CP Curve
Position [m]
0 0.5 1 1.5 2 2.5 3 3.5 4
-0.8

-0.6

-0.4

-0.2

0 Front Fairing
CP [-]

Rear Fairing
0.2

0.4

0.6

0.8

1.2

Figure 70 - Pressure coefficient plots of configuration 2

Once again one can refer to the pressure coefficient plots in Figure 70. The rear fairing shows a very steep

positive pressure gradient due to the sharper leading-edge. This would be favourable in laminar flow

regions. It also has a very steep pressure recovery region, but the pressure recovered is less than the

previous configuration. This means it will leave an even larger wake, which could be the reason for the

higher drag force. This makes it less favourable, adding to the reasons already mentioned.

Configuration 3

The third configuration had the same design for the front and the rear fairings, namely the NACA 16015-

YS900 combination fairings (Figure 71). These fairings had a drag force of 18.45 N, which, even though

it was slightly more than the first configuration discussed, was chosen as the configuration for the vehicle.

90
Figure 71 – Fairing Configuration 3

Figure 72 - Transition locations of flow on fairing Configuration 3

The reason for this is that the difference is marginal. Configuration 3 has a drag value of only 1.15% more

than Configuration 1. Furthermore, Configuration 3 has more advantages in that it simplifies the

manufacturing process. Instead of having to manufacture two moulds for the two different fairings, only

one mould has to be manufactured which could be used twice. Of the three configurations, the third

configuration has the smallest laminar flow regions on the rear fairing. The turbulent flow, however, also

has the lowest turbulent kinetic energy levels as can be observed in Figure 72. This is possibly why it only

shows a slight increase in the drag value.

91
Configuration 3 CP Curve
Position [m]
0 0.5 1 1.5 2 2.5 3 3.5 4
-0.6

-0.4

-0.2

0
Front Fairing
0.2
CP [-]

Rear Fairing

0.4

0.6

0.8

1.2

Figure 73 - Pressure coefficient plots of configuration 3

Referring to the pressure coefficient plots in the above figure (Figure 73), an assumption can be made that

the pressure recovered is the greatest of the three configurations. This also aids in reducing the overall drag

of the fairings and means that this configuration will have the smallest wake behind the rear fairing.

The frontal area of Configuration 1 and Configuration 2 is the same as the front fairing is the widest. There

will also be more space in the rear fairings of the third configuration for the suspension system and the

wheel and its braking mechanism. This third configuration was therefore selected for the preliminary detail

design and will be improved during the various revisions to follow.

7.1.5 Canopy Design

Three different canopy configurations were evaluated and compared. The different canopy configurations

were based on different symmetric airfoil profiles, as was done with the selection process of the fairings.

The canopy should allow enough space for the driver to have unrestricted movement. However, the sections

not utilized by the driver should not be too wide. This is to provide more space for the solar arrays.

92
According to the competition regulations the driver should be able to move 45° forward from a vertical

sitting position and 25° to the rear, without touching the canopy.

Canopy 1

The first canopy design was based on the design of the RUD20 airfoil. The aerodynamic properties of this

airfoil were discussed in the previous section and should therefore have favourable characteristics for the

design of the canopy.

When referring to the turbulent kinetic energy plot in Figure 74, it is clear that this canopy design shows

favourable laminar flow which extends to about 50% of the canopy length before transitioning to turbulent

flow. The laminar section at the bottom of the profile is disregarded because it does not have an influence

on the final selection criteria. This section will be cut from the canopy when integration with the vehicle

takes place. This canopy has an aerodynamic drag force of 10.83 N

Figure 74 - Canopy 1 turbulent kinetic energy plot

The pressure coefficient plots were analysed at about 50% of the final height of the canopy. This section

plane is indicated by the solid line in Figure 74. The same is done in the other canopy images. Analysing

the pressure coefficient plot of canopy 1 in Figure 75, the transition location is visible at the 2.5 m mark.

Note this is not the length of the canopy, but the position relative to the length of the entire vehicle, which

93
is 4.5 m. The canopy only has a favourable pressure gradient for about 40%. This is because the widest

section of the canopy is to the front, and the trailing section becomes narrower.

Canopy 1 CP Curve
Position [m]
1.6 1.7 1.8 1.9 2 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3 3.1 3.2 3.3 3.4
-0.5

-0.4

-0.3

-0.2

-0.1 Separation bubble


CP [-]

0.1

0.2

0.3

0.4

Figure 75 - Pressure coefficient plot of canopy 1

Regarding the space requirements for the driver, this canopy does meet the requirements. However, the

trailing section of the canopy recedes quite rapidly. This does not allow ample space for a roll bar that will

provide sufficient protection to the driver in the case of vehicle roll-over. The roll bar is an essential

requirement for the safety of the driver. For the reasons mentioned, this canopy will be modified to allow

more head room for the driver as well as more space for the roll bar integration in the trailing section.

Canopy 2

The second canopy design has a higher profile in the trailing section to allow for easier integration of the

roll bar. The canopy still allows the driver the required head space in the x-direction (forward and

backward) but the sideward movement of the driver is somewhat limited. This means a smaller frontal area

and a lower drag, but will limit the comfort of the driver. When the driver is fully kitted with a helmet, the

side to side movement will be limited which might cause driver fatigue over long distances.

94
Figure 76 - Canopy 2 turbulent kinetic energy plot

This design has better laminar flow than Canopy 1, extending beyond the halfway mark of the canopy to

between 60% and 70% of the chord length (Figure 76). This design also has a larger wetted surface so one

would assume it should have a higher drag force, but due to the laminar properties this design has a drag

force value of 9.15 N. This is a 15.51% improvement on the previous design.

The pressure coefficient plots of the canopy (Figure 77) were once again analysed at about 50% of the

height of the canopy. Canopy 2 has an improved pressure gradient over Canopy 1 and experiences this

positive gradient to about 60% before a negative gradient is experienced.

95
Canopy 2 CP Curve
Position [m]
1.6 1.7 1.8 1.9 2 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3 3.1 3.2 3.3 3.4
-0.4

-0.3

-0.2

-0.1
CP [-]

Separation bubble
0

0.1

0.2

0.3

Figure 77 - Pressure coefficient plot of canopy 2

The above plot shows that the flow transitions at about 75% chord length. This is a great improvement

over the previous canopy design. The canopy has a sharper pressure recovery region than Canopy 1 but

total pressure recovery is not achieved. To allow for more head room for the driver and a narrower trailing

section, this canopy design is modified further.

Canopy 3

Modifications to Canopy 2 resulted in a canopy with a narrower trailing section but still wide enough to

allow the driver ample head room as well as the integration of a roll bar. This also meant that Canopy 3

would have a very sharp pressure recovery region and an adverse pressure gradient. The second canopy

had sharp edges on the rear section but these have been replaced with thinner, rounded trailing edges.

96
Figure 78 - Canopy 3 turbulent kinetic energy plot

Canopy 3 does not have laminar flow that extends as far as the second design. This can be seen when

comparing the turbulent kinetic energy plots with one another (Figure 76 and Figure 78). This is partly due

to the larger frontal area, which is necessary for the comfort of the driver. The canopy is wider towards the

top of the canopy. This design has a drag force value of 9.44 N which is only a 3.07% increase over Canopy

2, but still 12.83% less than the first design.

97
Canopy 3 CP Curve
Position [-]
1.6 1.7 1.8 1.9 2 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3 3.1 3.2 3.3 3.4
-0.4

-0.3

-0.2

-0.1
CP [-]

0 Separation bubble

0.1

0.2

0.3

0.4

Figure 79 - Pressure coefficient plot of Canopy 3

Referring to Figure 79, a very sharp positive pressure gradient is experienced, transitioning to a gentle

gradient. The gradient continues to improve to about 65% chord length. The laminar flow transitions at

about 70% chord length after which a steep pressure recovery is experienced, as expected. Almost full

pressure recovery is experienced, which is an improvement over the previous canopy. This means that

Canopy 3 will have a smaller wake.

The larger frontal area has the effect of increasing the drag force, but as a better pressure recovery is

experienced, the drag increase from the previous design is not as great. The design has to be a trade-off

between aerodynamics and ergonomics, as driver comfort also plays a large role. Ergonomically, Canopy

3 is the better, design although it has a slightly higher drag force. Canopy 3 is therefore selected to be

integrated into the preliminary detail design of the solar-powered vehicle.

98
7.2 CFD Analysis of Detail Design

Effort was made to minimise the design parameters, especially the frontal area of the vehicle. The CAD

part was made using SolidWorks and then evaluated using STAR-CCM+, to determine the design with the

most favourable aerodynamic properties in terms of drag coefficient. Adjustments could be made to each

iteration, based on the drag coefficients and the transition locations determined when studying the turbulent

kinetic energy values on the surface of the model. These adjustments included variations in the fillet size

of the different junctions, the trailing edge thickness of the body and the appendages, as well as reduced

interference.

The detail design was completed by running simulations using STAR-CCM+ with the k-ω SST turbulence

model and the γ-Reθ transition model. In order to determine the number of cells where the accuracy

converged, namely the mesh independence, the simulation was run at different base size settings. The base

cell setting was started at 2500 mm and then reduced with each revision to determine the mesh

independence (Figure 80). It was assumed that the simulation reaches mesh independence at roughly 14

million trimmed cells. It was clear that there was no significant reduction in the drag force between 14

million and 17 million cells. The simulations were conducted at a base cells size of 250 mm which related

to about 14 million cells. The reason was that the computational power required was still manageable and

more accurate than with 8 million cells. Reducing the base size further would not result in a significant

increase in the accuracy of the simulation, but the increased computational power would have prolonged

the simulation significantly.

99
Mesh Independence
54

52

50
Force [N]

48

46

44

42

40
0.00E+00

2.50E+06

5.00E+06

7.50E+06

1.00E+07

1.25E+07

1.50E+07

1.75E+07

2.00E+07
Number of cells

Figure 80 – Mesh independence

The vehicle was simulated in a space of 32 m long, 15 m wide and 7 m high. This relates to 2 vehicle

lengths ahead of the model, 5 vehicle lengths behind the model and a tunnel which was more than 6 times

as wide and 5 times as high. This relates to the studies mentioned in the literature and ensures that the flow

ahead of the model is laminar. The trailing edge flow can also be properly captured as there is enough cells

behind the vehicle.

100
Symmetry Planes

Pressure Outlet

Vehicle Model
Velocity Inlet

Figure 81 - Detail design simulation boundary definitions

As discussed in the validation and verification chapters earlier, the inlet boundary of the simulated space

was treated as a velocity inlet set to a specific velocity. The simulations were set up as in the 2-D validation

case, where the model is kept stationary relative to the ground. This means that the air is simply blown past

the model. These results, discussed later, did not have the expected result. The simulations were therefore

re-done, to simulate the vehicle moving relative to the ground. This velocity was set to 28 m/s as in the 3-

D validation case. This relates to a Reynolds number of about 8.3x106 for a vehicle model with a length of

4.5 m. The outlet boundary was treated as a pressure outlet. The walls of the block were treated as

symmetry planes, therefore it was as though the vehicle was travelling in an open space on a level road.

The floor of the simulation domain was treated as a vector at the same tangential velocity as the inlet

boundary, in other words 28 m/s. This is to simulate the vehicle moving at a speed of about 100 km/h along

the tunnel and not simply standing still with the air blowing past it, at that same speed.

For a smooth transition from the vehicle body to the open space, the simulations were run using the

“Template Growth Rate” and the “Boundary Growth Rate” settings, set to “slow”. This ensures that the

cells are created evenly around the vehicle’s surface. As indicated, all simulations were conducted with a

base cell size of 250 mm, with refinements of 50 mm on the vehicle surface. Feature lines on the vehicle

101
are important for quality cells. The boundary layer was captured in 15 layers with a stretching factor of

1.5. The thickness of the boundary layer was set to 10 mm.

When analysing the results, the value for the drag force had the tendency of oscillating. These oscillations

are as a result of the flow of fluid over the body of the vehicle detaching from the surface of the vehicle,

thus becoming turbulent. These oscillations in the final drag force calculation can be very small, but it is

necessary to average the values of the last 100 iterations. It can be further reduced by using the polyhedral

meshing method, regarded to be more accurate, instead of the trimming method. A typical polyhedral cell

has 14 faces as opposed to the 6 faces of a hexahedral (trimmed) cell. This allows for a more accurate

solution. However, the number of cells created by the polyhedral mesher was too large when compared

with that of the trimmer and would increase the computing time substantially. The following sub-sections

describe the constraints that were adjusted with each simulation in order to reduce the drag coefficient of

the vehicle.

7.2.1. Model Based on Near Optimum Values

The first model which was based on near optimum values, according to the theory discussed, was based on

the same design as the baseline chosen during the conceptual design phase. The features of the model was

based on the theoretical optimum values as discussed in the literature, in order to reduce the design time of

the vehicle. The junction fillets in this case were 45 mm between the body and the fairings and 20 mm

between the body and the canopy. Fairing trailing edges had 7 mm fillets and the body trailing edge had a

thickness of 5 mm. The canopy had an L/h ratio of about 4. The trailing edge thickness of the parts would

not be varied as it complicated the simulation significantly in terms of obtaining a high quality mesh and

achieving convergence.

Applying the trimming method in STAR-CCM+ to the CAD model, the result was a simulation model with

a very fine mesh made up of about 13.5M cells. The boundary layer of the vehicle was captured by 15

prism layers, with a thickness of 10 mm. The drag force calculated was 40.7 N. With A in Eq. 2. 2 equal

to 0.789 m2, the CD value is calculated as 0.107.

102
The turbulent kinetic energy plots in Figure 82 shows where the laminar flow transitions to turbulent flow

on the body of the vehicle. This is where the colour scheme suddenly changes, almost representing a solid

line.

Transition

Figure 82 - Laminar flow on upper part of the baseline model

Figure 83 - Laminar flow on right side of the baseline model

Reviewing the turbulent kinetic energy plots of this baseline model (Figure 82 and Figure 83), it can be

estimated that the laminar flow on the upper part of the body extends to about 35% chord length. This

correlates with the individual analysis of the main body of the vehicle. This is not ideal, but the transition

position can be attributed to the thickness of the profile. A profile with a smaller frontal area might extend

103
the laminar flow, but this reduces space for internal components. The body profile is already at a minimum

thickness. The rounded edges of the leading-edge of the body do not cause transition.

The canopy shows favourable laminar flow, as already discussed in a previous section. The wake behind

the canopy, however, is quite large, but not much wider than the canopy thickness. This is expected. The

canopy can, however, be improved to reduce this wake and improve laminar flow. This might also aid in

the reduction of the high levels of turbulent flow on the body adjacent to the canopy, caused by the vortex

that is formed at the head of the canopy.

Vortices formed due to fairings

Small wake

Larger wake

Figure 84 - Laminar flow on lower part of the first vehicle model

Referring to the literature, it is stated that the optimum fairing junction fillet radius should be between 4-

6% of the chord length of said appendage. 45 mm is a good initiation point for the first revision. Referring

to Figure 84, it can be observed that the fairing junction fillets cause the flow to transition due to adverse

pressure gradient caused by the junction. This causes the laminar flow ahead of the fairing to transition.

The turbulent flow is then forced into a vortex along the side of the fairing. The transition of laminar flow

causes turbulent flow on the rest of the lower body section, with the farthest laminar flow experienced to

about 40% of the body.

Laminar flow on the front fairing is favourable, extending to about 80% of the chord length. This was

expected as the simulations of the fairing configurations showed this. The flow separates and then partially

re-attaches on the trailing edge of the fairings. Due to the re-attachment of flow and the strong pressure

104
recovery present during the individual simulation of the fairing configuration, the wake behind this fairing

is small. This still means that the rear fairing is caught in the turbulent wake. This turbulent wake causes

a larger vortex to be formed along the junction of the fairing and the body. Laminar flow, however, is still

experienced across about 50% of the surface of the rear fairing, but not as much as on the first fairing. This

is attributed to the vehicle moving at speed, ensuring that the flow can re-laminarise by the time it reaches

the rear fairing. Due to the larger vortex along the junction, a larger wake is experienced behind the rear

fairing. Laminar flow on the right hand fairing extends to about 40% chord length and partial reattachment

is also observed. The wake behind this fairly thick fairing is satisfactory. The sharper edges at the bottom

of the fairings also cause transition on the floor section of the fairings, on the inner side.

Systematic adjustments to be made to the current model are as follows:

 Adjust fairing junction fillets to see effect on transition of laminar flow and drag coefficient of the

vehicle. Fillet radii should theoretically be 4-6% of the chord length of the fairing.

 Adjust the canopy junction fillet radius.

 Left hand front fairing should be adjusted to allow rear fairing to experience more laminar flow.

 Increase fillet radius on the bottom edges of the fairings.

7.2.2. Effect of Increasing Fillet Radii on Fairings

The model that is described next remains the same with only the fairing junction fillets changing from 45

mm to 60 mm. This is to test the first point mentioned in the previous sub-section. The theoretical optimum

for the junction fillet radius is described as being 4-6% of the fairing length. The fairings being 1000 mm

long meant that the junction fillet would be between 40 mm and 60 mm. This revision will thus test if

enlarging the radius will allow the flow to progress smoother around the fairings. At this moment all fairing

radii are kept constant at 60 mm, regardless that the length of the right hand fairing is 3000 mm.

This resulted in the simulation model comprising 13.4M trimmed cells. The vehicle has A as 0.79 m2,

slightly larger than the previous model, due to the larger junction fillets. D is equal to 41.1 N, relating to a

CD value of 0.108. Comparing these results with that of the baseline model, there is a 1% and 0.1% increase

in the drag force and frontal area values, respectively. The value for CD slightly increases by 0.9%.

105
In the figure below (Figure 85) it is clear that the laminar flow does not change significantly by changing

the fillet radius to 60 mm.

Figure 85 - Laminar flow on upper surface of the model after changing the fillet radii to 60 mm

In Figure 86 the reason for the increase in the drag force is observed. The larger junction radius causes the

flow to transition earlier than in the previous section and therefore the horseshoe vortex is formed quicker.

This means that the laminar flow on the belly section of the vehicle does not extend quite as far. These

occurrences has the effect of increasing the drag force. The fillet radius is still within the optimum range,

so even though the fillet causes a larger wetted surface, the vortices created do not increase the drag

significantly as the increase is only about 1%.

106
Horseshoe vortex formed earlier

Figure 86 - Transition at fairing leading-edges of the revised model

The increase in the drag force and the drag coefficient is small and it can therefore be assumed that the 4-

6% fillet size theory still holds true. The fact remains that the increase in the radius increased the drag

force. It can be assumed that if this radius is increased further, especially on the shorter fairings, the increase

in wetted surface and frontal area will cause a much larger increase in the drag force.

The differences between the first two models of the vehicle are so little that either model can be used for

further optimisation. The following model will therefore serve as a test to see the effect on the drag

coefficient by reducing the radius (outside the theoretical optimum).

7.2.3. Effect of Reducing Fillet Radii to Below Theoretical Optimum

The third model revision of the vehicle, tests the fairing junction fillet radius outside the theoretical

optimum at 30 mm. This is 3% of the chord length of the fairing. For the theory tested in the previous two

revisions to hold true, this model should show an increase in the drag force value, even though there will

be a slight reduction in the frontal area of the vehicle.

The trimmed simulation model results in 13.5M cells, with A equal to 0.789 m2. This model had the same

drag force value as the first model. With the drag force equal to 40.4N, this is a reduction of 1.7% in the

drag force value from the model with the 60 mm fillet radius. The reduction, compared to the first model

with a 45 mm fillet, is negligible. The frontal area of this model is 0.2% smaller than that of the 60 mm

107
fillet radius model. The drag force measured on this model relates to a CD value of 0.107, which is a

reduction of about 1.5%, compared to the second model revision.

Figure 87 - Transition at fairing leading-edges of the 30 mm fillet radii model

One can see that this model has increased laminar flow at the leading-edge of the fairing junctions with the

body, when referring to Figure 87. It can also be observed that the effect of the vortex is not as great as

with the previous model, where a 60mm radius was implemented. The smaller radius has the effect of

reducing the effect of the vortex and thereby the total drag force of the vehicle. This relates to research

done by Maughmer et al. [41], where it was found that the smaller the fillet radius, the lower the drag force

on the junction. The reason for this is that the smaller fillet radius does not extend as far ahead of the fairing

as the larger fillet and the transition of flow is slightly delayed. The drag force is not increased greatly,

rather remaining the same as the 45mm fillet radius. It can therefore be assumed that the theoretical 4-6%

ratio for the fillet size is not a rule of thumb and a smaller fillet radius might be advantageous.

This same theory can be applied to the junction fillets of the canopy. The same procedure will be followed

as with the fairings. The fairing fillets will be reverted to the 45 mm size as it has a slightly larger frontal

area, but the same drag coefficient than the smaller 30 mm fillets. The canopy fillets will be changed from

the current 20 mm size to 45 mm fillets, to further test the theory.

108
7.2.4. Applying the Fillet Radius Theory to Canopy Junctions

Having tested the fairing junctions, the same method can be applied to the canopy junction with the body.

This is a more sensitive case as increasing the canopy junction radius, decreases space available for the

solar array. It is for this reason that the initial junction radius is chosen at 20 mm. The canopy junction

radius is set to 45 mm.

Adjusting the fillet radius from 20 mm to 45 mm resulted in a trimmed simulation model with 13.4M cells.

A is equal to 0.789 m2, showing that the larger fillet radius of the canopy junction had no significant effect

and the change is negligible (about 0.1%). The model was, however, still simulated in order to see the

effect that the larger radius will have on the overall drag force, as it is usually accepted that a smaller

junction will cause a horseshoe vortex, with a larger effect on the total drag force of the vehicle. The drag

force simulated was 40.5 N, about the same as the previous model, as well as the baseline model. At this

stage it can be assumed that the fillet radius will apply to the canopy junction fillet size as well. The drag

force and the frontal area results in a CD value of 0.107, as in the other similar models.

Reduced turbulent energy levels

Figure 88 - Canopy wake of the present model

When comparing Figure 88 with Figure 85, it shows that the increased canopy junction fillet radius has the

effect of slightly reducing the turbulent kinetic energy levels on the body adjacent to the canopy. This

might be attributed to the large radius, thus the effect of the vortex on the body is reduced, albeit only

slightly, with no significant change on the drag force.

109
7.2.5. Effect of a Large Canopy Junction Fillet

As described above, the canopy fillet radius is varied. The fifth model has a canopy fillet radius of 60 mm

and keeps the fairing fillets the same as in the previous revision. When the 60 mm fillet radius of the fairing

junction was tested in the second model, an increase of about 1% was experienced compared to the baseline

model with a 45 mm fillet. It can now be expected that the same trend will hold if the canopy junction fillet

is increased from 45 mm to 60 mm.

The simulation results in a vehicle with 12.6M trimmed cells. The value for A is 0.79 m2, which is 0.1%

greater than the previous model and about 0.15% greater than the first or baseline model. D is 41.1 N, 1.5%

greater than the model with the 45 mm canopy fillet and 1% in excess of the baseline model. This translates

to a CD value of 0.108, which is about 1.4 % greater than the previous revision and 0.8% greater than the

baseline model.

Taking this into consideration, it can be assumed that the same theory applies to the canopy junction fillets

as described for the fairing junctions, as the same trend of about 1% drag increase is experienced. It means

that when the vehicle is moving at a speed of about 100 km/h, the size of the junction fillets does not have

such a large effect on the overall drag of the vehicle, provided that it is within the theoretical optimum

tested here. The canopy junctions might also not have such a large effect on the drag force of the vehicle

because the canopy is already in a turbulent flow region. If the flow was laminar at the start of the canopy,

the vortex might have had a larger effect on the total drag force.

Looking at the revisions, it can be assumed that the best fillet radius size for the fairings, as well as for the

canopy is 45 mm and will be used in models to follow. It should be noted that this may not be the optimum,

but due to time constraints, more revisions are not conducted for the fillet radii. It just illustrates the method

that can be used to achieve a near optimum value. It can, however, be confirmed that the optimum radius

size is between 4-6% chord length.

7.2.6. Effect of Diverting Flow around the Rear Fairing

Having established the effects of fillet radii for this vehicle design, one can move to the third point as

outlined at the end of the first revision of this vehicle design. This is to adjust the front left hand fairing so

110
that the trailing flow will be channelled around the rear left hand fairing. This is achieved by using an

asymmetrical fairing; in this case, the same airfoil profile used to design the main body section. The 2-D

pressure coefficient for this profile was analysed and a favourable pressure gradient was present. The

fairing should therefore experience good laminar flow, reducing the overall drag force of the vehicle. The

fairing should allow the flow to pass on the inner side of the rear fairing. The fairing fillets are kept at 45

mm.

The simulation results in a vehicle with 13.4M trimmed cells. The value for A is 0.794 m2, which is roughly

0.5% greater than the previous model (canopy fillet adjustment) and 0.6% greater than the baseline model.

This larger increase is due to the new fairing being wider at the top, where it is integrated with the body.

This change in the geometry sees the drag force reducing to 39.5N, 3.8% less than the previous model and

2.9% less than the baseline. This translates to a CD value of 0.104.

Increased Laminar Flow

Figure 89 - Laminar flow on the sixth model

In Figure 89 the improved laminar flow on the rear left hand fairing is clear. The turbulent flow on the

lower section of the rear fairing near the ground, slightly increases the drag force, but not as much as when

the rear fairing is completely in the turbulent wake of the front fairing. This increased laminar flow on the

rear fairing, ensures that even with an increase in the frontal area of the vehicle, there is actually a decrease

experienced in the drag coefficient of this model. There is more turbulent flow adjacent the fairing on the

inner side of the body, but this side of the fairings also sees a significant increase in the laminar flow.

111
Figure 90 - Laminar flow on inside of fairings

There are still high levels of turbulent flow on the lower part of the rear fairing (Figure 90), which might

be improved if the fairing does not have such sharp edges at the bottom. This will be improved on in a later

model revision. The wake behind the fairings is also clear in Figure 91.

Effect of modified front fairing, albeit the


larger wake

Figure 91 - Fairing wake due to the modified front fairing

This fairing configuration, different from the configurations discussed in the previous section, actually

creates a larger wake than two identical fairings, as can be seen when referring to Figure 91. However, the

increased laminar flow on the rear fairing ensures that the larger wake does not result in an increase in the

drag force. The increased turbulent flow around the rear fairing, as mentioned, is visible in red.

It can be assumed that the advantages of the configuration discussed here, outweighs that of the two

identical fairings. Even though it complicates the manufacturing process, the increased laminar flow may

increase the efficiency of the vehicle in race circumstances.


112
7.2.7. Applying Fillet Radius Theory to the Driver Cockpit Fairing

Having established that the two different fairings give a positive change, the following step is to once again

change the theoretical optimum for the fairing fillet radius. It was mentioned previously that the length of

the long faring, which includes the driver cockpit, has been ignored until a later revision. This fairing has

a chord length of 3000 mm, and if the same theory is to be applied, the fairing fillet radius should be set to

between 120 mm (4% of chord) and 180 mm (6%). For the shorter fairings a value of 4.5% was used and

therefore the long fairing fillet radius will be set to 135 mm. Once again it is assumed that the theoretical

optimum value will hold true and a reduction in the drag coefficient will be experienced.

The simulation results in a vehicle with 12.4M trimmed cells. The value for A is 0.797m2, which is roughly

0.4% greater than the previous model without the 135mm fairing fillets and 1% greater than the baseline

model. This increase is due to the 135 mm fillet radius used on the long fairing. This change in the

geometry sees the drag force increase to 40.0 N, 1.2% greater than the fairing modification, but 1.7% less

than the baseline model. The drag force translates to a CD value of 0.104, which is equivalent to the previous

model. The model experienced an increase in the frontal area, and this together with only a slight increase

in the drag force (0.5N), compared to the previous model, results in the drag coefficient remaining the same.

Increased Fillet Radius

Figure 92 – Increase in frontal area due to larger fillets on the right hand fairing (left side in above image)

Figure 92 shows the larger fillet of the right hand fairing with the associated increase in the frontal area of

the vehicle. As mentioned this increase does not result in an increase in the drag coefficient of the vehicle.

Once again it can be ascribed to the reduction of the effect of the vortex created at the start of the fairing

113
and it can be assumed that for this model the theoretical optimum for the fairing junction size once again

holds true. This theoretical optimum should be investigated with all future vehicle designs, as it is not a

rule that it will reduce the drag force. It is largely dependent on the effects of the vortex created by the

junction, which in some cases might be reduced by a larger fairing radius.

7.2.8. Effect of Removing Sharp Floor Edges on Fairings

The next step to improve on the drag coefficient is the final point mentioned at the end of the baseline

model section. This is to increase the radius of the bottom edges of the fairings to promote a smooth

transition of flow from the leading-edge of the fairing to the bottom floor section.

This results in a simulation with 12.7M trimmed cells. The value for A is 0.801 m2, which is roughly 0.5%

greater than the previous model and 1.5% greater than baseline model. This larger increase is due to the

new fairing being slightly wider due to the changed geometric nature of the fairing design. This change in

the geometry sees the drag force reduced to 39.1 N, 2.3% less than the second to last model (effect of

modifying the front fairing) and 4.0% less than the baseline model. This translates to a CD value of 0.102,

which is a reduction of 2.8% lower the previous model with the larger fairing. This drag coefficient is also

a 5.8% reduction from the baseline model. This satisfies the objective to achieve a reduction of at least 5%

in the drag coefficient of the vehicle.

Rounded edges on fairings

Figure 93 - Rounded lower edges of fairings

Increasing the radius of the lower edge of the fairings can be observed in the above figure, Figure 93. This

change also sees an increase in the overall length of the fairings, albeit slightly. The fairing can be scaled

114
down in the design phase in order to maintain the same chord length as in previous models, but this will

also reduce the length of the lower section of the fairing. This, in turn, means that the width of the fairing

will either have to remain the same (thus scaling the fairing in the x direction only and thereby elongating

it), which will alter the airfoil properties, possibly reducing the laminar flow, or scaling the fairing

uniformly meaning that the width of the fairing is reduced and thereby the space available for the wheels

and suspension. The fairing is thus elongated.

Figure 94 - Improved laminar flow on lower edge of fairings

The increase in the radius of the lower edge does achieve what is expected, namely to increase the smooth

transition of laminar flow from the leading-edge of the fairing to the lower section, without causing

transition to turbulent flow prematurely (Figure 94).

Figure 95 - Improved laminar flow on rear left fairing

115
The same flow properties is not experienced on the rear fairing, probably because of the turbulent wake

caused by the front wheel (Figure 95). Even though this is the case, the drag force of the vehicle is not

increased, and the drag coefficient is further improved.

7.3 Detail Design Results

It was mentioned that the simulations were done with the model stationary relative to the ground. These

results are summarised in Table 6. The result of the last model increased, and as it was expected to decrease,

it was opted to redo the simulations with the vehicle moving relative to the ground.

Table 6 - Drag force results of the stationary vehicle

Model 1 Model 2 Model 3 Model 4 Model 5 Model 6 Model 7 Model 8


F 43.0 43.0 40.4 42.1 43.8 42.3 41.4 41.9
CdA 0.089 0.090 0.091 0.088 0.091 0.088 0.086 0.087
A 0.789 0.790 0.789 0.789 0.790 0.794 0.797 0.801
Cd 0.113 0.113 0.115 0.111 0.116 0.111 0.108 0.109

The step where the fillets were optimised for both the long and short fairings (Model 7, Table 6) had the

lowest drag coefficient value of 0.108, which is 4.8% lower than the very first model simulated and before

any changes to the features of the vehicle were made. The last model of the vehicle saw an increase of

0.5% compared to this model, which was unexpected.

The results of the variation in the design of the solar-powered vehicle, when moving relative to the ground,

is summarised in Table 7. It can clearly be seen in Figure 96 how the drag coefficient varies with the

changes to the design, as described in the preceding section of the dissertation.

Table 7 - Drag force results of the moving vehicle

Model 1 Model 2 Model 3 Model 4 Model 5 Model 6 Model 7 Model 8


F 40.7 41.1 40.4 40.5 41.1 39.5 40.0 39.1
CdA 0.085 0.086 0.084 0.084 0.086 0.082 0.083 0.081
A 0.789 0.790 0.789 0.789 0.790 0.794 0.797 0.801
Cd 0.107 0.108 0.107 0.107 0.108 0.104 0.104 0.102

116
Drag Coefficient of Models
0.109

0.108

0.107
Drag Coefficient [-]

0.106

0.105

0.104

0.103

0.102

0.101
1

8
Model

Figure 96 - Drag coefficients of the vehicle model when moving relative to the ground

The last model, with the rounded fairing edges and the optimised fillets for both the long and short fairings

(Section 7.2.8), had the lowest drag coefficient value of 0.102, which is 5.8% lower than the very first

model simulated and before any changes to the features of the vehicle was made. Reviewing the above

results, it should be noted that the vehicle should always be simulated when moving relative to the ground,

instead of being stationary.

7.4 Conclusion

These iterations and modifications to the vehicle show that it is possible to improve the vehicle design by

following the methodology described. It was stated in the objectives that a reduction of 5% in the drag

coefficient is deemed significant, provided that a near optimum base model is used. The results showed

that the base model was improved by almost 6%. It should be noted that this was achieved when the vehicle

was simulated to be moving relative to the ground rather than remaining stationary. This is a more realistic

representation of reality and causes a significant reduction in the drag force calculated. It can therefore be

assumed that this is a suitable methodology, although many more iterations and modifications may be

117
required to significantly reduce the drag of the vehicle. By applying this methodology to future designs,

this is possible. Reducing the effect of vortices created by the appendage fillets, by keeping the fillets

within a theoretical optimum and in some cases reducing the fillet size below this theoretical optimum

value, is the most significant contribution in the reduction of the drag coefficient of the vehicle. Adding to

this, it should be ensured that all the appendages experience laminar flow, meaning that none of the

appendages should be in the turbulent slip-stream of another appendage. The methodology also proves that

CFD is indeed a valuable tool in the design process and should be continually implemented and improved

upon. It significantly reduces the design time required and the need for the use of wind tunnels. These

results can be further validated by the use of a wind tunnel to test the as-built design of the vehicle and

comparing this with the CFD results.

118
Chapter 8: Conclusion and Recommendations

The main objective of this study is to optimise the aerodynamics of a solar-powered vehicle using CFD.

This had to be conducted within the Sasol Solar Challenge regulations. In doing this, the method is

described that can be used to analyse and design future vehicles and it is proven that CFD is a viable tool

with which to conduct the work. Lastly, the transition models used in the CFD code had to be validated.

The conclusions that are made from the study are summarised under the headings below.

8.1. 2-D Validation

The NACA 0018 airfoil was used for the 2-D validation, as well as a trimmed mesh setting. The airfoil

was tested at different angles of attack, which were achieved by using an overset mesh function within

STAR-CCM+. This correlates with the experimental tests conducted by Timmer which was also conducted

at a range of angles of attack. At lower angles of attack, the CFD results compared well (within 5%) with

these experimental results, as well as with the results obtained from the XFoil-based XFLR5. At higher

angles of attack, these values differed by more than 10%. CFD simulations were conducted using the γ-

Reθ transition model coupled with the k-ω SST turbulence model. Lift coefficients were compared with

coefficients of drag, as well as lift coefficients at different angles of attack. It can be concluded that using

the transition models for a 2-D simulation achieves satisfactory results.

8.2. 3-D Validation

The 3-D validation was conducted on the HPT’s VeloX1 human-powered bicycle. The CFD simulation

results were compared with the wind tunnel results conducted on the bicycle by the HPT from the Delft

University. Furthermore, results conducted by the same institution using VSAero, were also compared with

the CFD results.

As with the 2-D validation, the simulation was conducted using the γ-Reθ transition model coupled with

the k-ω SST turbulence model. The transition locations determined by CFD showed similarities to the wind

tunnel results, extending to about halfway the body length, but differed in other respects. The value for the

drag area (supplied by the Delft University) was larger than that calculated by the CFD. Reasons for this

may be attributed to differences in the geometry, such as surface roughness and irregularities due to
119
manufacturing. The actual model also has ventilation ducts, wheel holes and a windscreen, all contributing

factors to higher drag values.

As transition position is a better way to determine the accuracy of the simulations, rather than using the

drag area or drag force value, it was assumed that the CFD will predict the laminar flow accurately enough

to be used for the purposes of the detail design of the solar-powered vehicle. It would seem that the

transition model have been satisfactorily validated for the 3-D simulations of the vehicle.

8.3. Detail Design

This iteration-based methodology whereby the initial design parameters are modified, proved to work well

when multiple revisions of the vehicle is used. Analysing each revision and then modifying the design

according to the results and the available literature, enabled one to determine the positive or negative

changes to the design. This methodology can thus be assumed to be effective in the design of solar-powered

vehicles.

The results indicated that CFD is indeed a valuable tool to be used in the design process and is an essential

part of the detail design phase as it reduces the need for wind tunnels. It also reduces the design time

required as each iteration of the design can be evaluated within a few hours if the resources are available.

The results also indicated the importance of using the correct simulation settings. If the vehicle is simulated

to be stationary relative to the ground, the calculated drag force will be significantly higher than when the

vehicle is simulated to actually move relative to the ground.

8.4. Recommendations for Future Work

Future work can be divided into the different components as analysed in this study as well as the overall

design. The main body of the vehicle should be further optimised to extend the laminar flow past the 50%

mark, as opposed to the current 35% mark. As indicated in the revisions, the front fairing should be

designed such that the rear fairing experiences more laminar flow. The front design should, however, be

further analysed as increased laminar flow is possible. A turbulent airfoil for the rear fairing is also an

alternative to the current design. The canopy size can be decreased, thus optimised for size and shape,

which would further reduce the overall drag of the vehicle.

120
With the integration of the components, effort should be made to reduce all junctions. This means, aligning

the outside edges of the fairing with that of the outside edge of the main body. In this way the junctions

are reduced considerably. The leading fillet of the fairing should also be modified to have a gentler slope

(in excess of the current 45°) as described in the literature. The trailing edges of the fairings should be

sharper. As indicated in the last model, the bottom of the fairings should be rounded. Wheel wells can also

be implemented to channel the air that is sucked into the body through the wheel holes to the outside of the

body. The overall manufacturing process of the car should be improved to ensure that body panels have

smaller seam gaps. It should also be noted that any change in the surface geometry can cause transition of

flow. Stickers used for branding should thus be avoided. Any branding on the vehicle should be painted

onto the surface of the body to ensure a smooth finish.

Taking into account the above, all of these factors can be optimized simultaneously. This will eliminate

the trial and error methodology of adjusting a single variable at a time, as described in this dissertation.

This means that a mathematical algorithm can be developed, which will use a number of input variables

and vary these parameters simultaneously in order to determine the best possible combination. These

parameters are then applied to the model which can be simulated to determine the drag coefficient of the

vehicle.

Experimental tests should be developed in order to validate the CFD code with the actual as-built design of

the vehicle. It was assumed that the CFD code was accurate to such a degree that it could be used for these

simulations. This assumption is based on the 3-D validation process conducted in this dissertation.

121
Bibliography

[1] D. W. Kurtz, Aerodynamic design of electric and hybrid vehicles: a guidebook, Pasedena: Jet

Propulsion Laboratory, 1980.

[2] CD-adapco, Dynamics:Simulating Systems Issue 33, Melville, NY: CD-adapco, 2013.

[3] S. i. Sport, “Sasol Solar Challenge,” 2014. [Online]. Available:

http://www.solarchallenge.org.za/?page_id=468. [Accessed 17 September 2014].

[4] Wikipedia, “Tour de Sol,” 2012. [Online]. Available: http://en.wikipedia.org/wiki/Tour_de_Sol.

[Accessed 28 February 2013].

[5] G. Tamai, The Leading Edge: Aerodynamic Design of Ultra-streamlined Land Vehicles, 1st ed.,

Cambridge, USA: Bentley Publishers, 1999.

[6] H. Smith, The illustrated guide to aerodynamics, Blue Ridge Summit, PA: Tab Books, 1992.

[7] A. Watts, “Semi-truck Boat Tail Improves Fuel Efficiency 7.5%,” 5 November 2009. [Online].

Available: http://wattsupwiththat.com/2009/11/05/semi-truck-boat-tail-improves-fuel-efficiency-7-

5/. [Accessed 8 April 2013].

[8] “Aerospaceweb.org,” 1997-2012. [Online]. Available:

http://www.aerospaceweb.org/question/aerodynamics/q0215.shtml. [Accessed 8 April 2013].

[9] A. Shapiro, Shape and Flow: The Fluid Dynamics of Drag, New York: Anchor Books, 1961.

[10] R. Beardmore, “Roymech,” 28 January 2013. [Online]. Available:

http://www.roymech.co.uk/Related/Fluids/Fluids_Drag.html. [Accessed 09 March 2014].

[11] S. R. Ahmed, B. Bayer, J. Callister, N. Deuβen, H.-J. Emmelmann, H. Flegl, A. George, A. Gilhaus,

H. Götz, H. Groβmann, R. Hoffmann, W.-H. Hucho, D. Hummel, W. Kohl, G. Mayr, G. Necati, R.

122
Piatek, D. Schlenz, W. Sebbeβe, N. Singer and P. Steinberg, Aerodynamics of road vehicles, 4th

ed., W. Hucho, Ed., Warrendale, Pa.: Society of Automotive Engineers, 1998.

[12] VATSIM, “VATSIM,” 2014. [Online]. Available:

http://www.vatnz.net/cms/index.php?option=com_content&task=view&id=181&Itemid=173.

[Accessed 9 March 2014].

[13] A. Filippone, Advanced Aircraft Flight Performance, New York: Cambridge University Press,

2012.

[14] “Cortana Corporation,” [Online]. Available: http://www.cortana.com/Drag_Description.htm.

[Accessed 9 April 2013].

[15] R. B. Langtry, “A Correlation-based Transition Model using Local Variables for Unstructured

Parallelized CFD Codes,” Institut für Thermische Strömungsmaschinen und

Maschinenlaboratorium: Universität Stuttgart, Stuttgart, 2006.

[16] B. R. Munson, D. F. Young, T. H. Okiishi and W. W. Huebsch, Fundamentals of Fluid Mechanics,

John Wiley & Sons (Asia) Pte Ltd, 2010.

[17] “The Free Dictionary,” 2013. [Online]. Available:

http://encyclopedia2.thefreedictionary.com/Separation+of+Flow. [Accessed 11 April 2013].

[18] H. Schlichting and K. Gersten, Boundary-layer Theory, Berlin: Springer-Verlag, 2000.

[19] F. Menter, R. Langtry and S. Volker, “Transition Modelling for General Purpose CFD Codes,”

Flow Turbulence Combust, vol. 77, pp. 277-303, 18 August 2006.

[20] M. Germano, U. Piomelli, P. Moin and W. H. Cabot, “A dynamic subgridscale eddy viscosity

model,” Physics of Fluids, vol. 3, no. 7, pp. 1760-1765, July 1991.

[21] K. Walters and D. Cokljat, “A Three-Equation Eddy-Viscosity Model for Reynolds-Averaged

Navier-Stokes Simulations of Transitional Flow,” Journal of Fluids Engineering, vol. 130, pp.

121401.1-101401.14, December 2008.

123
[22] F. R. Menter, T. Esch and S. Kubacki, “Transition Modelling Based on Local Variables,”

Enginnering Turbulence Modelling and Experiments, vol. 5, pp. 555-564, 16-18 September 2002.

[23] R. B. Langtry and F. R. Menter, “Transition modeling for General CFD Applications in

Aeronautics,” in AIAA Aerospace Sciences Meeting and Exhibit, Reno, Nevada, USA, 10-13

January 2005.

[24] F. R. Menter, T. Esch and S. Kubacki, “Transition modelling based on local variables,” in 5th

International Symposium on Turbulence Modeling and Measurements, Mallorca, Spain, 2002.

[25] F. R. Menter, R. B. Langtry and S. R. Likki, “A Correlation-Based Transition Model Using Local

Variables - Part 1: Model Formulation,” Journal of Turbomachinery, vol. 128, no. 3, pp. 413-422,

2006.

[26] R. B. Langtry, F. R. Menter, S. R. Likki, Y. B. Suzen, P. G. Haung and S. Volker, “A Correlation-

Based Transition Model Using Local Variables - Part 2: Test Cases and Industrial Applications,”

Journal of Turbomachinery, vol. 128, no. 3, pp. 423-434, July 2006.

[27] S. Mansor, N. A. R. Nik Mohd and C. W. Chung, “Validation of CFD Modeling and Simulation of

a Simplified Automotive Model,” Applied Mechanics and Materials, vol. 735, pp. 319-325, 2015.

[28] V. K. Yakkundi and S. S. Mantha, “Computational Analysis of “Drag & Lift” of a car With K-ε std,

and K-ω SST (Shear Stress Transport) Turbulence Models & Effect of Mesh Refinement,” Curie,

vol. 3, no. 3 and 4, pp. 117-129, 2010-2011.

[29] D. R. Caroll, The Winning Solar Car, Warrendale: SAE International, 2003.

[30] G. Buresti, G. V. Iungo and G. Lombardi, “Methods for the drag reduction of bluff bodies and their

application to heavy road-vehicles,” DDIA, vol. 6, p. 4, 2007.

[31] B. W. McCormick, “Static Stability and Control,” in Aerodynamics, Aeronautics, and Flight

Mechanics, John Wiley & Sons, Inc., 1995, pp. 473-483.

124
[32] “Longitudinal Static Stability,” Stanford, 04 January 2005. [Online]. Available:

http://adg.stanford.edu/aa241/stability/staticstability.html. [Accessed 21 October 2014].

[33] “Chapter 7: Fixed-Wing Aerodynamics and Performance,” Train Army, 27 12 2012. [Online].

Available: https://rdl.train.army.mil/catalog-ws/view/100.ATSC/124A2C90-72CD-4B05-AB8B-

ABC8718F9760-1274574464617/3-04.203/chap7.htm. [Accessed 21 October 2014].

[34] K. E. Kenyon, “Curvature Boundary Layer,” Physics Essays, vol. 16, no. 1, pp. 74-85, 2003.

[35] S. Ruitenberg, “S&D Tech Blog,” 14 February 2015. [Online]. Available:

http://sdtechblog.weebly.com/uploads/6/5/2/5/6525158/7411858_orig.jpg. [Accessed 26 February

2015].

[36] L. M. M. Boermans, F. Nicolosi and K. Kubrynski, “Aerodynamic Design of High-performance

Sailplane Wing-fuselage Combinations,” in 21st ICAS Conference, Melbourne, 13-18 September

1998.

[37] R. L. Simpson, “Junction Flows,” Annual Review of Fluid Mechanics, vol. 33, pp. 415-443, 2001.

[38] A. Ahmed and M. J. Khan, “Effect of sweep on wing-body juncture flows,” in 33rd American

Institute of Aeronautics and Astronautics Aerospace Sciences Meeting, Reno, 1995.

[39] S. F. Hoerner, Fluid-Dynamic Drag, Vancouver: Published by author, 1965.

[40] S. F. Hoerner, Fluid-dynamic drag : practical information on aerodynamic drag and hydrodynamic

resistance, Bakersfield, California: Hoerner Fluid Dynamics, 1992.

[41] M. Maughmer, D. Hallman, R. Ruszkowski, G. Chappel and I. Waitz, “Experimental Investigation

of Wing/Fuselage Integration Geometries,” Journal of Aircraft, vol. 26, no. 8, 1989.

[42] “World Solar Challenge,” [Online]. Available:

http://www.worldsolarchallenge.org/files/136_solar_team_twente.large.jpg. [Accessed 2015

February 2015].

125
[43] “Nuon Solar Team,” 2012. [Online]. Available: http://www.nuonsolarteam.nl/?lang=en. [Accessed

20 March 2012].

[44] N. S. Team, 8 August 2013. [Online]. Available:

http://www.flickr.com/photos/nuonsolarteam/9814081074/. [Accessed 23 January 2014].

[45] J. Haringman, “solarracing.org,” 2013. [Online]. Available:

http://www.solarracing.org/2013/08/29/tokai-university-reveals-their-car-challenger/. [Accessed 11

March 2014].

[46] T. Abe and P. Starr, “Evolving an Engineering Design Methodology,” in International Conference

of Engineering Education, Budapest, 2008.

[47] D. G. Ullman, The Mechanical Design Process, New York: McGraw-Hill, 1992.

[48] B. E. Kissinger, “Aerodynamic optimization of a solar race car body shape,” UMI, Ann Arbor,

1997.

[49] AAIA, “Guide for the verification and validation of computational fluid dynamic simulations,”

American Institute of Aeronautics and Astronautics, Vols. AIAA-G-077-1998, 1998.

[50] J. R. Grace and F. Taghipour, “Verification and validation of CFD models and dynamic similarity,”

Elsevier, vol. 139, no. Powder Technology, pp. 99-110, 2004.

[51] “Profili 2,” 2015. [Online]. Available: http://www.profili2.com/eng/default.asp. [Accessed 2015

February 23].

[52] W. A. Timmer, “Two-dimensional low-Reynolds number wind tunnel results for airfoil NACA

0018,” Wind Engineering, vol. 32, no. 6, pp. 525-537, 2008.

[53] CD-adapco, “The world's most comprehensive engineering simulation inside a single integrated

package,” 2013. [Online]. Available: http://www.cd-adapco.com/products/star-ccm-plus. [Accessed

14 5 2013].

126
[54] K. Maley, “Best Practices: Volume Meshing [PowerPoint],” CD-adapco, 2012.

[55] N. E. Ahmad, E. Abo-Serie and A. Gaylard, “Mesh Optimization for Ground Vehicle

Aerodynamics,” vol. 2, no. 1, 2010.

[56] U. o. Strathclyde, “Sustainable Engineering,” [Online]. Available:

http://www.esru.strath.ac.uk/EandE/Web_sites/09-10/MCT/html/Technical/rfoil.html. [Accessed

2015 February 23].

[57] K. Epema, Interviewee, [Interview]. 26 June 2013.

[58] B. Maskew, “Program VSAERO Theory Document,” National Aeronautics and Space

Administration, Redmond, Washington, 1987.

127
Annexure A: 2-D Validation Summary Report

This section lists the settings used in STAR-CCM+ to simulate a NACA 0018 airfoil on a 2-D mesh. All

the settings described in the respective chapter of this dissertation correlates with the settings described in

this section.

128
129
130
131
132
133
134
135
136
137
Annexure B: 3-D Validation Summary Report

As in the previous section, the settings listed in this summary report correlates with the simulation settings

used in STAR-CCM+ to simulate the flow around a 3-D body, the VeloX1 human powered bicycle.

138
139
140
141
142
143
144
145
146
147
Annexure C: Detail Design Analysis Summary Report

The following summary report lists the STAR-CCM+ settings used during the 3-D evaluation of the solar

powered vehicle in question. It is based on the settings of the previous section.

148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
Annexure D: Final Remarks

Due to time constraints, the vehicle built to compete in the Sasol Solar Challenge, was the baseline model

as described in Section 7.2.1. The moulds for the vehicle were constructed exactly as defined by the CAD

part of the model. These moulds were split so that the different body panels would fit exactly onto each

other. During the manufacturing process of the vehicle, several implications became apparent. These

included that the solar panels would not fit onto the top of the car, so the split line of the top body panel

had to be adjusted. The bypass diodes were installed onto the wrong solar panels, so the bubble canopy of

the vehicle was modified to be smaller. This resulted in a poorer canopy design, possibly resulting in higher

drag, but the frontal area was decreased. This effect was simulated and the poorer canopy design had a

0.6% in drag force. This is negligible so the effects of the poorer aerodynamic shape and the reduction in

the frontal area eliminated one another almost completely.

Figure 97 - NWU Solar Car competing in the Sasol Solar Challenge 2014

In Figure 97 the modified bubble canopy design can be observed. This is one of the main components that

requires a redesign and the design should not only be modified. Further implication with the rear suspension

169
meant that the vehicle ride height was approximately 50 mm higher than initially designed for. This would

also increase the drag force of the vehicle.

The centre of gravity was theoretically determined to be 617 mm behind the centre of the front wheels and

864 mm left of the left-hand wheel track. It is also 250 mm above ground level. This puts the static margin

at almost 10% in front of the centre of the vehicle. This ensured that the vehicle was extremely stable on

the road. When the vehicle was driven around the Zwartkops Raceway in Centurion, Pretoria, the top speed

achieved was 130 km/h. This could be achieved without the vehicle experiencing any traction loss on the

wheels and is testament to the stability of the aerodynamic design.

The results of the vehicle during the race is summarised as follows:

 Final over-all distance covered by the team in the vehicle was 2360 km. This puts the team in an

over-all 4th place and second of the South African teams that participated.

 Record for the furthest distance travelled by a South African team in a single day was 416 km.

Finally, it should be noted that the vehicle would have performed better if more time was available to test

the vehicle. If faults with the electrical system, such as the faulty battery pack, had been discovered earlier,

it could have been corrected in time for the start of the race.

Below follows more photographs of the vehicle during the Sasol Solar Challenge 2014.

170
Figure 98 - Sirius X25 headed to Kroonstad on Day 1 of the Sasol Solar Challenge 2014

Figure 99 – Sirius X25 travelling on the N1 highway between Kroonstad and Bloemfontein, Free State.

171
Figure 100 – On route to Hanover in the Karoo

Figure 101 - Sirius X25 completing a loop in the Karoo

172
Figure 102 - The vehicle on its way to Port Elizabeth, Eastern Cape

Figure 103 - Doing a loop between Heidelberg and Witsand in the Western Cape

173
Figure 104 - Control stop in Heidelberg, Western Cape

Figure 105 - Sirius X25 and team at the finish in Cape Town

174

You might also like