You are on page 1of 16

The current issue and full text archive of this journal is available at

www.emeraldinsight.com/0961-5539.htm

Effects of rear spoilers on ground


vehicle aerodynamic drag

Effects of
rear spoilers

Halil Sadettin Hamut and Rami Salah El-Emam


Faculty of Engineering and Applied Science,
University of Ontario Institute of Technology (UOIT), Oshawa, Canada

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

Murat Aydin
Faculty of Engineering and Applied Science,
University of Ontario Institute of Technology (UOIT), Oshawa, Canada and
Institute of Energy, Istanbul Technical University, Istanbul, Turkey, and

627
Received 20 March 2012
Revised 20 September 2012
Accepted 12 October 2012

Ibrahim Dincer
Faculty of Engineering and Applied Science,
University of Ontario Institute of Technology (UOIT), Oshawa, Canada
Abstract
Purpose The purpose of this paper is to examine the aerodynamic effects of rear spoiler geometry
on a sports car. Today, due to economical, safety and even environmental concerns, vehicle aerodynamics
play a much more significant role in design considerations and rear spoilers play a major role in this area.
Design/methodology/approach A 2-D vehicle geometry of a race car is created and solved using
the computational fluid dynamics (CFD) solver FLUENT version 6.3. The aerodynamic effects are
analyzed under various vehicle speeds with and without a rear spoiler. The main results are compared
to a wind tunnel experiment conducted with 1/18 replica of a Nascar.
Findings By the CFD analysis, the drag coefficient without the spoiler is calculated to be 0.31.
When the spoiler is added to the geometry, the drag coefficient increases to 0.36. The computational
results with the spoiler are compared with the experimental data, and a good agreement is obtained
within a 5.8 percent error band. The uncertainty associated with the experimental results of the drag
coefficient is calculated to be 6.1 percent for the wind tunnel testing. The sources of discrepancies
between the experimental and numerical results are identified and potential improvements on the
model and experiments are provided in the paper. Furthermore, in the CFD model, it is found that
the addition of the spoiler caused a decrease in the lift coefficient from 0.26 to 0.05.
Originality/value This paper examines the effects of rear spoiler geometry on vehicle aerodynamic
drag by comparing the CFD analysis with wind tunnel experimentation and conducting an uncertainty
analysis to assess the reliability of the obtained results.
Keywords Aerodynamic drag, Drag and lift coefficient, Rear spoiler, Wind tunnel
Paper type Research paper

Nomenclature
A area (m2)
b
CD
CL
FD
FL
L

standard systematic uncertainty


drag coefficient
lift coefficient
drag force (N)
lift force (N)
length (m)

M
P
r
Re
s
U
V

number of measured data points


pressure (kPa)
experimental results
Reynolds number
standard deviation
overall uncertainty
velocity (m/s)

The authors would like to thank Dr Akif Ezan for all his help and support with modeling the
vehicle geometry and UOIT Engineering Lab Technician Qi Shi for helping them with the
experiments.

International Journal of Numerical


Methods for Heat & Fluid Flow
Vol. 24 No. 3, 2014
pp. 627-642
r Emerald Group Publishing Limited
0961-5539
DOI 10.1108/HFF-03-2012-0068

HFF
24,3

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

628

Xi measurements of the ith variable


Greek symbols
r density (kg/m3)

m
tij
P

viscosity (kg/m-s)
shear-stress tensor (N/m2)
dependent variable

1. Introduction
Today, due to economical, safety and even environmental concerns, vehicle
aerodynamics play a much more significant role in design considerations than it did
before. It is estimated that the aerodynamic drag is the governing form of resistance
when vehicles reach speeds of 80 km/h or greater, especially considering the fact
that 65 percent of the power required at 110 km/h is consumed due to overcoming
aerodynamic drag (Leduc, 2009; Diamond, 2004). Therefore the improvements in
aerodynamic characteristics can result in significant decrease in driving stability,
handling, fuel consumption and overall efficiency. Thus, it is no surprise that rear
spoilers are becoming more widely used in the auto industry. With increasing in the oil
prices and more stringent legal regulations, more research is conducted on accessories
that can improve the vehicles fuel efficiency. Among these accessories, rear spoilers
are one of the major ones due to being relatively inexpensive, easy to install and
aesthetically appealing. Furthermore, it can reduce the drag coefficient and increase
the fuel efficiency significantly for most passenger vehicles, while decreasing the lift
coefficient in the expense of the drag coefficient in most race cars. However, the benefits
of the rear spoiler are not limited only to fuel cost. By also deflecting the traveling air
upwards by increasing the pressure rear deck, the rear spoiler can provide better
traction, faster turning, acceleration and brake as well as increasing the vehicle safety.
Even though the rear spoiler is one of the key players in vehicle aerodynamics, there
are only limited amount of associated studies in the literate that incorporates both
the computational analysis and experimental verification. The studied vehicles vary
from mini-vans to modified race cars and are analyzed for various purposes such as
improving aerodynamic and aero-acoustic characteristics, preventing exhaust smoke
patterns and even reducing the dirt and/or snow accumulating on the rear surface of
the vehicle (Kiyoshi et al., 1997; Parihar et al., 2006).
Tsai et al. (2009) analyzed the effects of five different rear spoilers on the aerodynamics
and aero-acuostics of a Honda S2000 under 180 km/h. The authors constructed the model
in ICEM/CFD and used FLUENT as the CFD solver with k-A model. They have
determined the sound pressure levels for each case and also calculated drag coefficients
ranging from 0.4 to 0.51 and lift coefficients ranging from !0.001 to 0.06. The lowest drag
coefficient is calculated in the case without the spoiler and it increases in different spoiler
designs in the attempts of reducing the lift coefficient. Zake (2008) analyzed BLM car
with spoiler and touring wings in order to understand their aerodynamic effects on the
vehicle between 70 and 150 km/h and determined that the lift coefficient is reduced in
the expense of the drag coefficient. Kim (2004) studied the effects of rear spoilers on wake
flow characteristics and drag for large-sized buses. The author assumed symmetry and
created half of the 3-D model of the vehicle from the centerline, without the side mirrors.
The author used RNG k-A model in order to incorporate the secondary straining effect
that is not considered with the standard k-A model and set the Reynolds number to
5.27 " 106 based on the body height and inlet velocity. The author calculated the drag
coefficient for the vehicle to be 0.518, which is within 4 percent of the experimental value
determined in a wind tunnel by the 1/16-scale of a model commercial bus in the reference.
By adding a rear spoiler to the vehicle, the initial extreme vertex produced on the
cross-car axis at the rear upper body was removed, in the expense of creating a new

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

vertex on the rear end surface and reduced the drag coefficient by more than 12 percent.
Kim et al. (2008) also developed a rear spoiler for a mini-van and conducted analysis using
CFD standard and RNG k-A models at Reynolds number 2.72 " 106 with and without
the spoiler and found reduction in drag and lift by 5 and 100 percent, respectively.
As can be seen from the literature, there is a significant lack of studies that
examines the effects of rear spoilers on vehicles, that combines computational analysis
and experimental testing. Moreover, the authors do not consider the error associated
with the uncertainties throughout the experimental process. Thus, the purpose of
this study is to examine the aerodynamic effects of rear spoiler geometry in on a race
car under various wind speeds using a CFD solver FLUENT. The results are compared
to the wind tunnel testing of a scaled down vehicle model and uncertainty analysis is
conducted to assess the reliability of the results.

Effects of
rear spoilers

629

2. Experimental setup
The 1/18-scale Nascar model was tested in a Flotek wind tunnel designed by the GDJ
Inc as seen in Figure 1. The wind tunnel has exceptionally consistent velocity
profiles across the test section, with turbulence measures of less than 0.2 percent, and

(4,202 mm)
TOTAL LENGTH
(3,977 mm)
STAND LENGTH
(861 mm)
FRONT STAND
WIDTH
(1,067 mm)
FRONT TUNNEL
WIDTH

(731 mm)
REAR STAND
WIDTH

(3,847 mm)
TUNNEL LENGTH

Steel Contraction cone

Steel Diffuser Section


Clear Acrylic Test Section
Mounted on 1.5 in. Steel Angel

(1,600 mm)
TOTAL HEIGHT

PVC Shelf
Control Box Mount
1.25 x 1.25
Steel Supports

Source: Per GDJ Inc.

Fan Guard

2 x 2 Steel Frame

2 HP DC Motor

Figure 1.
Flotek 1440 wind tunnel
used in the experiment
and its dimensions

HFF
24,3

630

allows speeds up to 40 m/s. It has a 4200 " 4200 inlet and 140 overall length with 1200 "
1200 " 3600 visible test section.
For the experiment, a 1/18-scale model of a Nascar is used with nine pitot tubes
connected across the top body geometry in the same plane (except for the additional
spoiler measurement), in order to get the velocity and pressure distribution of the wind
flowing over the vehicle. A strain gauge is connected at the bottom in order to calculate
the drag force of the vehicle under the velocities tested. The pressure/velocity and
drag force measurements are monitored and calculated through a software, while the
pressure measurements are also calculated through manometers on the wind tunnel.
The scaled model with the sensors can be seen in Figure 2.

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

2.1 Buckingham-Pi verification


The wind tunnel experiment was conducted using a 1/18-scale of a Nascar in a wind
tunnel with 4200 " 4200 inlet and 140 overall length. In order to confirm that the scaled
model would be a good representation of the actual prototype, dimensionless analysis
is conducted, ensuring that the geometric, kinematic and dynamic similarities are
achieved. To be able to make this comparison, Reynolds number is selected as the
independent dimensionless parameter and a variation of the standard drag coefficient
is selected as the dependent parameter as seen in Equation (3) (compressibility
parameters are neglected due to low wind velocities, Mach number is o0.1). BuckinghamPi theory reveals that when the chosen independent variable (Reynolds number) between
the model and the prototype is determined to be the same, then the dependent variable in
the prototype (drag coefficient) is guaranteed to be equal to the dependent variable in the
model (Cengel and Cimbala, 2006). Thus, the drag coefficient for the actual vehicle can be
determined as long as the Reynolds number between the model and the prototype are
determined to be similar:
FD
rVL
P1 f P2 ; where P1
and P2
1
rV 2 L2
m
Here, air density for the measured wind tunnel temperature is determined to be
1.184 kg/m3 along with a dynamic viscosity of 1.849 " 10!5 kg/m-s. The velocity of
the actual prototype is taken as 30 m/s and the characteristic length of 2.4 m is used
6

4
3

Figure 2.
1/18-scale Nascar
model along with 13
measurement points

2
1

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

for the car. The actual length is used for the characteristic length based on the same
assumption used in the vast majority of the studies conducted (Macciacchera and Ruck,
2001; Caldichoury et al., 2011). With the aforementioned parameters, the Reynolds
number is calculated to be 4.6 " 106.
If the temperatures are taken to be the same for the prototype and model testing,
then the density and viscosity parameters will be similar for both cases. Hence the
significant reduction is size (1/18 scale) requires the wind velocity to be 18 times higher
to achieve similarity. Since the prototype speed is usually already 30 m/s, the model is
required to be exposed to 540 m/s, which is not feasible with the wind tunnel used.
In order to work around this limitation, correlation between drag coefficient and
Reynolds number is developed for practical use based on the experimental data. From
Figure 4, it can be seen that around Reynolds number of 2.9 " 104, the drag coefficient
becomes independent of the Reynolds number. By using this correlation, it can be assumed
that the drag coefficient will not differ significantly between the highest achieved wind
velocity and the required wind velocity to achieve dynamic symmetry. Therefore drag
coefficient value is calculated using the aerodynamic drag force at this velocity.
2.2 Experiment
In order to verify the accuracy of the Fluent model and compare it to real life case, the
experiment is conducted in the wind tunnel. Special attention has been given to have
to match the CFD model parameters (such as temperature and wind speed) and well as
making it as realistic as possible. The vehicle is placed in the middle of the test area of
the wind tunnel facing the incoming wind and secured tightly to reduce any potential
discrepancy in measurements. The data for the pressure and velocities on the vehicle
upper body based on different wind speeds are retrieved.
The vehicle grill (in point 1, Figure 3) has the highest pressure and zero velocity
since it is the stagnation point of the flow. As the wind travels over the vehicle, its
pressure initially is reduced, causing an increase in the velocity, and increases again
as it finally reached to the rear of the vehicle, reducing its associated velocity, but still
being less than the initial due to frictional forces. The comparison of pressure and
velocity distributions is illustrated in Figure 3.
The measurements from the strain gauge are displayed in the drag force section
of the software and provided the respective force in grams force. By knowing the
associated air density, vehicle frontal area and wind velocity, the drag coefficient is
calculated using the drag force equation:
1
FD rV 2 ACD
2
2
Based on the air density at 251C and frontal area of 31.5 cm2, the drag coefficient of the
vehicle can be calculated with respect to any tested wind velocity. The correlation
between the wind velocity and drag coefficient can be seen in Figure 4. The points
are best fitted with respect to available data.
It should be noted that, however, there will be some additional error associated
with the experimental calculations. The source of this error comes from the fact that
the cross-sectional area of the wind tunnel is relatively small, which can have
significant impact on the flow velocities. If the cross section of the wind tunnel is not
large enough, the effective freestream wind passing the vehicle can accelerate
significantly which can increase the aerodynamic drag and therefore prevent the
similarity conditions to be achieved. In order to determine the impact of this effect,

Effects of
rear spoilers

631

HFF
24,3

Top Surface Pressure


centimeters of water
Grill 0.07
Hood 3.89

0
25

Grill
Hood

24

Hood
Hood
Roof
Roof

40

10

30

15

20

23
26
26

Trunk 3.17

20

10

23

Trunk

Trunk 0.78

25

11

Trunk

20

Spoiler

Spoiler 2.49

8 7

Converging
accelerating flow
(7)

8 7

Low Pressure
High Speed
(5-6)

3 2

Converging
accelerating flow
(2-4)

High Pressure
Low Speed
(1)

High Pressure
Low Speed
(8-9)

Figure 3.
Pressure and velocity
distribution over
the vehicle
0.38
y = 2E10x 3 2E06x 2 + 0.0041x 2.9816

0.375

R 2 = 0.97

0.37
0.365
0.36
CD

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

Hood 3.20
Roof 4.26
Roof 4.10

Hood 3.52

632

Top Surface Velocity


(m/s)
50

0.355
0.35
0.345
0.34

Figure 4.
Drag coefficient
as a function of
Reynolds number

0.335
0.33
2

2.2

2.4

2.6

2.8

Re(104)

wind tunnel blockage can be calculated based on the ratios of the model frontal area
and the cross-sectional area of the wind tunnel test section.
The scaled model frontal area is measured to be 0.00315 m2 and the manufacturers
tunnel specifications is used for the respective cross section (0.0929 m2), creating a

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

blockage ratio of 3.3 percent. Since the acceptable blockage ratio in the literature
is determined to be up to 7.5 percent (Holmes, 2001), this value will not have much
significant impact on the drag coefficient determined by the experiment. Furthermore,
the flow profile in the wind tunnel will be slightly different than the actual scenario due
to the vehicle being stationary during testing. This will result in a boundary later build
up on the wind tunnel floor, which changes the underbody flow from the actual case
and hence prevent kinematic similarity.
The lift coefficient on the other hand (shown in Equation (3)) was not been able to be
calculated due to the limitations of the experimental setup. Because of the space
constraints inside the model vehicle, the pitot tubes were placed only on the upper
portion of the vehicle, therefore aerodynamic effects of the underbody were not been
able to be captured properly. Moreover, the available software was not capable of
calculating the lift force associated with the vehicle. Hence only drag force measurements
are used for the analysis:
FL 2rV 2 AFl

Effects of
rear spoilers

633

3. Uncertainty analysis
In order to assess the reliability of the results from the wind tunnel experiments,
one should consider the error associated with systematic and random uncertainties
throughout the experimental process. For the study, single-sample experiment
uncertainty will be analyzed due to relatively low number of independent data points
taken at each test point as well as the having relatively small fixed error. In order to
determine the overall uncertainty of the experiment, the contributors of uncertainty
associated with the capability of the instrument and the instability of the processes are
merged with the help of Taylor series expansion and root-sum-square methods.
Initially the fixed and variable errors associated with the wind tunnel system
are determined by not varying the experimental process (zeroth-order level). Next, the
process instability and random instrument error in the process is determined running
different processes using the same equipment, procedures and instrumentation
(first-order level). Finally the overall uncertainty in the experiments, including the
effects of process unsteadiness, is estimated by taking the root-sum square of the fixed
errors due to instrumentation and first-order uncertainty (Nth-order level) (Moffat, 1988).
3.1 Zeroth-order analysis
The drag coefficient of the vehicle is determined from the data obtained through the
experiments using the modified version of Equation (4). The uncertainties associated
to each term in the equation are determined with respect to the measurement errors for
each element. The zeroth-order estimates of the systematic and random standard
uncertainties for the aforementioned variables can be seen in Table I.
Variable
FD
r
V
A

Value

Systematic standard uncertainty

Random standard uncertainty

1.2152 Na
1.184 kg/m3
35.41 m/s
0.0035 m2

0.004 kg/m3

0.004 m2

0.02 N

0.3 m/s

Note: aDrag force with respect to the highest wind velocity of 35.41 m/s

Table I.
Zeroth-order estimates of
systematic and random
standard uncertainties for
variable in drag coefficient
determination

HFF
24,3

634

The uncertainties in the measured variables cause the uncertainty in the results (r in
this case) and hence is often modeled using a propagation equation based on Taylor
series expansion to find the systematic standard uncertainty (br) is given below
(Coleman and Steele, 2009):
"
"
J !
J !
X
#
$
qr 2 2 X qr 2 2
4
bxi
sxi
for r r X1 ; X2 . . . XJ
b2r
qXi
qXi
i1
i1

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

where bxi is the standard systematic uncertainty and sxi is the standard deviation for
measurement of each variable Xi. The partial derivatives needed for the above
equation are given below:
qCD
2

qFD rV 2 A

qCD
!2FD
2 2
qr
rV A

qCD
6FD

qV
rV 3 A

qCD
2FD

qA
rV 2 A2

By substituting the partial derivatives into the Taylor series expansion equation, the
systematic uncertainty (with respect to air density and vehicle cross-sectional area) is
calculated as follows:
!
"2 !
"2
qCD
qCD
2
bA
br
9
bCD
qr
qA

which results in bCD 0:0014 after substituting the values in the above equation.
By using the same method, the random uncertainty associated with the expression
(with respect to drag force and velocity) is given below:
!
"2 !
qCD
qCD 2
bv
b FD
10
s2CD
qFD
qV
which results in sCD 0:0205 after substituting the values in the above equation.
Thus the large-sample expression for the combined standard uncertainty UCD, with
a 95 percent confidence, can be calculated as shown in the equation below:
%
&12
UCD 2 b2CD s2CD 0:0412

11

This uncertainty, which shows the suitability of the wind tunnel, corresponds to 12.1
percent of the calculated drag coefficient.
3.2 First-order analysis
By having ten repeated measurement values of the variable CD, the sCD value is known
from previous experiments as 0.02. If the first-order random standard uncertainty is

divided into two categories, namely zeroth-order and sample-to-sample variation, the
random standard uncertainty due to sample-to-sample variation can be calculated as
given below:
h# $
i12
# $
# $2
2
sCD SAMPLE VAR sCD 1ST ! sCD ZEROTH 0:021
12

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

This uncertainty, which estimates the scatter in the results of different trials,
corresponds to 6.9 percent of the calculated drag coefficient.

Effects of
rear spoilers

635

3.3 Nth-order analysis


Finally the overall uncertainty of the mean result (with an uncertainty if 95 percent) is
obtained from the following equation:
!
"12
sC
2
2
UCD 2 bCD s
0:0192
where sCD pffiffiDffiffiffi 0:0095
13
CD
M

where the bar symbol above the drag coefficient represents the mean value and M is
the number of measured data points. This uncertainty, which represents the overall
uncertainty composed of fixed errors due to instrumentation and first-order
uncertainty, corresponds to 6.1 percent of the calculated drag coefficient.

4. Numerical analysis
When a vehicle is driving on the road, as the air flows through the vehicle, it moves
towards the point A in Figure 5 (so called stagnation point), where the static and total
pressures become equal. From this point, the flow divides into, above and below the
vehicle. Determining the flow pattern underneath the vehicle can be very complex and
depends on various factors such as the height of the underbody and the presence of
fairings. Thus raising the vehicle underbody and/or having smooth fairings that cover
the mechanical elements underneath the vehicle can lower the air drag significantly.
However, these options are usually ignored when compared against other issues this
may causes such as more costly maintenance (Genta, 1997).
In point B of Figure 5, the flow velocity increases by the pressure being less the
total (or even less than the ambient) pressure. Between points C and D, the flow
detached and attached back again due to the sharp change in the vehicle geometry.
The pressure distribution between points E and F primarily depends on the shape of
the vehicle roof, but will still be relatively low. At the end of the roof the velocity slows

D
B

Source: Adapted from Genta (1997)

Figure 5.
Streamlines around a
passenger vehicle in the
symmetry plane

HFF
24,3

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

636

down as the pressure rises and the flow detaches after point F. This phenomenon can
be seen in Figure 5.
This phenomenon is demonstrated by cutting the vehicle geometry in rubber and
placing it in a Hele-Shaw-type fluid visualization apparatus, as well as placing the
actual scaled replica in the wind tunnel with a piece of string on the upper body, which
in both cases, give an idea on the boundary streamlines, separation and wake regions
in the rear portion of the vehicle. The initial one is done with the help of water and
red dye between two glass sheets where the vehicle geometry is placed downstream of
the water flow. Even though the streamlines look non-ideal due to the limitations of the
experiment, the flow separation in the rear portion of the vehicle and respective wake
regions can be clearly seen, as shown in Figure 6.
The latter one is demonstrated in the wind tunnel where the flow of the string
guides as streamlines over the vehicle. It can be seen that, with the rear spoiler, the flow
separation occurred further behind the vehicle which affects the drag force at this
location. This is shown in Figure 7.
4.1 Governing equations
The external flow around the car can be considered laminar or turbulent according to the
speed of the car. Two-dimensional incompressible steady-state Navier-Stokes equations
were used to predict the associated aerodynamic features. The continuity and momentum
equations, that are seen below, are solved numerically with k-A turbulence model:
qu qv
0
qx qy
!

qu
qu
r u v
qx
qy

"

qP
q 2 u q2 u

!
m
qx
qx2 qy2

14
!

!
!
"
qv
qv
qP
q2 v q2 v
!
m

r u v
qx
qy
qy
qx2 qy2

Figure 6.
Streamlines over the
vehicle geometry in the
Hele-Shaw-type fluid
visualization apparatus

15a
15b

Effects of
rear spoilers

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

637

Figure 7.
The streamline
representation with
respect to the string
placed over the vehicle
in the wind tunnel

The standard k-A model is a semi-empirical model that is based on model transport
equations for the turbulence kinetic energy, k which have been driven from the exact
equation, and the other transport equation represents the dissipation rate and it was
obtained from the physical reasoning and bears little resemblance to its mathematically
exact counterpart. This model has been used extensively during the last two decades
and has been widely accepted for engineering applications. While the standard k-A
turbulence model was being the focus of turbulence modeling, another turbulence
modeling techniques were also intensively developed by researchers. May be some of
them are more superior and more user-friendly for general CFD users than the standard
k-A turbulence model. However, the standard k-A turbulence model performs better as
far as the computational time and cost are considered. In this study, standard k-A
turbulence model is chosen to predict the flow field around the vehicle.
4.2 CFD analysis
The computational domain of the model is prepared on ICEM Ansys software. The
model represents the external flow over a 2-D model of the car which resides in a wind
tunnel. Two different cases are studied for the same car geometry with and without spoiler.
The car model length is 2.56 m and its height is 0.7 m. Also the computational domain
representing the tunnel walls are considered as 13 m long and 5 m high. The mesh is
created for the whole computational domain. For model a, Figure 8(a) which represents the
model of the car without spoiler, an orthogonal mesh is created, while for the second case,
with the spoiler is meshed using tri-mesh trend, Figure 8(b) which is recommended for
meshing models dealing with aerodynamic studies. The meshes are refined until mesh
independence is reached; mesh refining is performed in the area around the car where the
flow characteristics are predicted to be affected significantly. Table II shows the mesh
fining steps that are taken. The computational mesh is shown in Figure 8. The uniform
velocity is specified as the inlet boundary condition, outlet flow is specified for pressure
outlet. The top and down boundaries of the computational domain are selected as
wall where we are modeling the flow of the air around the car body in the wind tunnel.

HFF
24,3

638

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

Without spoiler

Figure 8.
The computational
meshes used for the car
without and with the
spoiler

With spoiler

Number of cells

Table II.
Mesh independency

26,058
75,474
120,436

Drag coeff. (without spoiler)

Number of cells

Drag coeff. (with spoiler)

0.29
0.31
0.31

36,841
84,345

0.36
0.36

Inlet velocity applied as 30 m/s (Re 1 " 106). The drag coefficient and lift coefficient are
dimensionless forms of drag and lift. They are defined with respect to Equations (4) and
(5), where the area A is defined as the characteristic area of the body, which is considered
as the frontal area (the projected area seen when looking to the object in a direction parallel
to the upstream velocity) for the drag coefficient calculations. For the lift coefficient
calculations, the planform area (which is represented by the projected area observed
by looking towards the object in a direction normal to the upstream flow) is considered.
4.3 Numerical results
Finite volume method is used through which the conservation principles were
applied to the model. The governing differential equations are integrated to yeild a set
of algebric equations to ensure all the quantities are conserved. Next, these algebric
equations are solved through numerical means to obtain the unknown quantities.
The standard SIMPLE scheme was utilized to solve the pressure velocity coupling
discretized equation. First-order upwind scheme was used for discretization of
the turbulence kinetic energy and the turbulence dissipation rate. The solver uses
under relaxation to control the update of computed values at each iteration. The under
relaxation factor applied to this study were 0.1, 0.3, 0.8 and 0.8 for pressure, momentum,
turbulence kinetic energy, turbulence dissipation rate, respectively.
The drag coefficient calculated for the car without the spoiler as 0.31 and for the case
with spoiler 0.36 while the lift coefficient decreased from 0.26 to 0.05. The spoilers are
used for decreasing the drag coefficient, however, for most sports cars; the lift force is

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

more important as it controls the car stability, thus the drag coefficient can be sacrificed
in order to achieve the desired lift coefficient. Figure 9 represents the shape of the stream
lines of the external flow around the car body with and without the spoiler at the same
operating conditions. It is seen that the spoiler caused a deflection in the streamlines near
the car body which affected its shape in the rear area after the car body.
The static pressure contours are represented for both with and without spoiler cases
in Figure 10, as it is clear that the pressure over the rearend of the car with the spoiler is
higher, which is going to increase the drag force across the spoiler and hence for the
body of the vehicle. It can be also concluded that a sudden deceleration of the flow
occurs in the front bumper which creates an increase in the static pressure.
As calculated by the experiments, the front part of the vehicle is simulated to have
zero velocity and highest pressure, shown both through the CFD analysis and wind
tunnel experiment (Figures 10 and 11, along with experimental point number 1 in
Figure 3, respectively). Subsequently, the flow accelerates in a rapid trend over the top
of the body causing a pressure suction zone, which is indicated by the blue region in
simulation results provided in Figure 10, along with the high-velocity region in the
experimental points 5 and 6. Finally, the pressure gradually increased as the flow
velocity decreased and the flow eventually detached from the vehicle, which is also
captured through these figures.
Even though the CFD model creates significant value in terms of predicting the
aerodynamic drag and lift associated with the vehicle under predetermined conditions,
verifies the experiment outcomes and the associated literature values within small
error margin; there are some shortcomings/potential improvements associated with the
model. Due to the complicated geometry and unavailability of data of the tested vehicles
exterior geometry, the model is based on a more generic passenger vehicle geometry.
Furthermore, in order to save computational time, the model is constructed in 2-D
which was unable to consider some of the real life effects associated with the cross-car
air flow distribution. It should be expected that, due to having no runoff area for the air
flow that flow over or under the vehicle, some of the flow that travels from the top or
bottom of the vehicle would actually separate from the side and travel through the
vehicle side. Since this phenomenon is not taking into consideration, the acceleration
over and under the body could be over predicted, resulting in larger velocities in these
regions. Moreover, the 2-D model did not incorporate the wheels in the geometry which

Effects of
rear spoilers

639

Without spoiler

With spoiler

Figure 9.
Stream lines of the air
flow around the car body

HFF
24,3

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

640

6.48e+02
5.69e+02
4.90e=02
4.10e=02
3.31e+02
2.51e+02
1.72e+02
9.23e+01
1.28e+01
6.66e+01
1.46e+02
2.25e+02
3.05e+02
3.84e+02
4.64e+02
5.43e+02
6.23e+02
7.02e+02
7.82e+02
8.61e+02
9.41e+02

Without spoiler

Figure 10.
Contours of the static
pressure (Pa)

6.19e+02
5.23e+02
4.26e+02
3.29e+02
2.32e+02
1.35e+02
3.84e+01
5.85e+01
1.55e+02
2.52e+02
3.49e+02
4.46e+02
5.43e+02
6.40e+02
7.36e+02
8.33e+02
9.30e+02
1.03e+03
1.12e+03
1.22e+03
1.32e+03

With spoiler

reduces the drag force calculated in the simulation, due to reducing the frontal area of
the vehicle. Finally, the separate flows inside the actual vehicles, namely the cooling
flow inside the engine department and the passenger cabin, are also ignored in the
model. Even though the passenger cabin effects are usually negligible, the air flowing
through the engine department can affect the overall drag coefficient significantly
(Launder and Spalding, 1974; Theera-apisakkul and Kittichaikarn, 2005).
5. Conclusions
This paper examines the aerodynamic effects of rear spoiler geometry on a race car. 2-D
vehicle geometry is constructed using ICEM commercial package and the results are
compared with a wind tunnel experiment of a 1/16 scale of the actual vehicle.
Through the CFD analysis, the drag coefficient without the spoiler is calculated to be
0.31. When the spoiler is added to the geometry, the drag coefficient increases to 0.36.

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

Effects of
rear spoilers

4.38e+01
4.23e+01
4.09e+01
3.94e+01
3.79e+01
3.65e+01
3.50e+01
3.36e+01
3.21e+01
3.06e+01
2.92e+01
2.77e+01
2.63e+01
2.48e+01
2.33e+01
2.19e+01
2.04e+01
1.90e+01
1.75e+01
1.61e+01
1.46e+01
1.31e+01
1.17e+01
1.02e+01
8.76e+00
7.30e+00
5.84e+00
4.38e+00
2.92e+00
1.46e+00
0.00e+00

641

Without spoiler

4.82e+01
4.58e+01
4.34e+01
4.10e+01
3.86e+01
3.61e+01
3.37e+01
3.13e+01
2.89e+01
2.65e+01
2.41e+01
2.17e+01
1.93e+01
1.69e+01
1.45e+01
1.20e+01
9.64e+00
7.23e+00
4.82e+00
2.41e+00
0.00e+00

Figure 11.
Velocity contours (m/s)
With spoiler

The computational results with the spoiler are compared with the experimental data,
and a good agreement is obtained within a 5.8 percent error band. The sources of errors
are identified along and potential improvements on the model and experiments are
provided in the paper. Furthermore, in the CFD model, it is found that the addition
of the spoiler caused a decrease in the lift coefficient from 0.26 to 0.05.
References
Caldichoury, I., Delpin, F. and Lapoujade, V. (2011), Incompressible CFD results using LS-DYNA.
For high Reynolds number flow around bluff bodies, Eighth European LS-DYNA Users
Conference, Strasbourg, May.
Cengel, Y.A. and Cimbala, J.M. (2006), Fluid Mechanics: Fundamentals and Applications, ISBN
0072472367, The McGraw-Hill Companies Inc., New York, NY.

HFF
24,3

Downloaded by UNIVERSITY OF SOUTHAMPTON At 07:02 17 March 2015 (PT)

642

Coleman, H.W. and Steele, W.G. (2009), Experimentation, Validation, and Uncertainty for
Engineers, 3rd ed., John Wiley & Sons Inc, Hoboken, NJ.
Diamond, S. (2004), Heavy vehicle systems optimization, annual progress report for heavy
vehicle systems optimization, Washington, DC.
Genta, G. (1997), Motor Vehicle Dynamics: Modeling and Simulation, ISBN 10: 9810229119, World
Scientific Publishing Co., Hackensack, NJ.
Holmes, J. (2001), Wind Loading of Structures, 2nd ed., Taylor & Francis, London.
Kim, M. (2004), Numerical study on the wake flow characteristics and drag reduction of
large-sized bus using rear-spoiler, International Journal of Vehicle Design, Vol. 34 No. 3,
pp. 203-217.
Kim, K., Geng and, X. and Chen, C. (2008), Development of a rear spoiler of a new type of
minivans, Journal of Vehicle Design, Vol. 48 Nos 1/2, pp. 114-131.
Kiyoshi, Y., Hase, N., Fujita, S., Isomura, R., Takeda, I., Sumitani, K. and Murayama, T. (1997),
Concurrent CFD Analysis for Development of Rear Spoiler for Hatchback Vehicles, Society
of Automotive Engineers Inc, Warrendale, PA.
Launder, B.E. and Spalding, D.B. (1974), The numerical computations of turbulent flows,
Computer Methods in Applied Mechanics and Engineering, Vol. 3 No. 2, pp. 269-289.
Leduc, G. (2009), Longer And Heavier Vehicles, An Overview Of Technical Aspects, ISBN 978-9279-12983-4, JRC Scientific and Technical Reports, European Communities, Luxembourg.
Macciacchera, I.U. and Ruck, B. (2001), Pressure fluctuations induced by road vehicles in
ambient air a model study, Proceedings of the Workshops on physical modelling of
environmental flow and dispersion, University of Hamburg. Hamburg, September 3-5.
Moffat, R.J. (1988), Describing the uncertainties in experimental results, Experimental Thermal
and Fluid Science, Vol. 1 No. 1, pp. 3-17.
Parihar, A., Kulkarni, A., Stern, F., Xing, T. and Moeykens, S. (2006), Using FlowLab, an
educational computational fluid dynamics tool, to perform a comparative study of turbulence
models, Computational Fluid Dynamics Journal, Vol. 15 No. 1, pp. 175-182.
Theera-apisakkul, K. and Kittichaikarn, C. (2005), Numerical Analysis of Flow Over Car Spoiler,
Computational Mechanical Laboratory, Department of Mechanical Engineering, Faculty of
Engineering, Kasetsart University, Bangkok.
Tsai, C., Fu, L., Tai, C., Huang, Y. and Leong, J. (2009), Computational aero-acoustic analysis
of a passenger car with a real spoiler, Applied Mathematical Modelling, Vol. 33 No. 9,
pp. 3661-3673.
Zake, R. (2008), Aerodynamics of aftermarket rear spoiler, BS thesis, University of Malaysia, Pahang.
Corresponding author
Halil Sadettin Hamut can be contacted at: Halil.Hamut@uoit.ca

To purchase reprints of this article please e-mail: reprints@emeraldinsight.com


Or visit our web site for further details: www.emeraldinsight.com/reprints

You might also like