You are on page 1of 4

The Blockage Effects on Vehicle Aerodynamics in Closed-Wall

Wind Tunnel A CFD Study


Prasanjit Das
Candidate for the Degree of Master of Engineering
Supervisor: Professor Makoto Tsubokura
Division of Mechanical and Space Engineering

Abstract
An investigation into the total blockage effects on
aerodynamic drag in closed-wall wind tunnel testing
has been carried out by computational fluid dynamics
(CFD). The simulation of flow past two different
realistic, full-scale, heavy-duty truck models was
conducted. The models were placed at different yawing
angles in the virtual wind tunnels with blockage ratio of
about 10%; which is exactly the same as the DNW
experimental wind tunnel geometry. To compare the
results with ideal (blockage free condition) case, other
simulations with larger cross section at the blockage
ratio of 1.1% were also conducted. Comparisons of the
ideal CFD results to the corrected Cd values obtained
by some existing blockage correction methods for
closed-wall wind tunnel are also examined.
Key words: Wind tunnel blockage Corrections,
Aerodynamic drag, Closed-wall wind tunnel, CFD

Introduction
At the development stage, it is important to assess the
vehicle performance in account of engine power. In a
typical class of heavy-duty truck, power required to
overcome rolling resistance and accessories increases
linearly with vehicle speed, while energy losses due to
aerodynamic drag increase with the cube of the speed.
At a typical highway speed of 70 mph, aerodynamic
drag accounts for approximately 65% of the energy
output of the engine. Even modest reductions in
aerodynamic drag can significantly reduce fuel
consumption. Lower fuel consumption will result in a
reduction in pollution emissions, and, more
importantly, a reduce dependence on fossil fuels. In
automotive industry, to assess the performance of road
vehicle, aerodynamic testing is normally conducted in
wind tunnel. For the constrained effects imposed by the
walls of a wind tunnel around the vehicle, there is
always a need to correct the aerodynamic quantities
such as aerodynamic forces and moments. The
corrected results can then reflect the performance under
blockage free conditions. To date, most of the blockage

correction equations of closed-wall wind tunnel were


based on simplified bluff body geometry. Cooper et al.
[1] compared a number of correction methods and
recommended not to use the classical formulae for
vehicle aerodynamics. Wind tunnel corrections for
separated flows were developed first by Maskell [2].
His analysis was originally derived for circular plates
normal to the flow and it based on conservation of
momentum. Recently, a semi-empirical correction was
proposed by Mercker [3], which for the first time
introduced additional blockage effects associated with
the pressure gradient along the test section at empty
condition. The purpose of the present study is to
investigate the closed wall wind tunnel blockage effect
on full-scale heavy-duty truck by using CFD, to assess
the blockage correction equations for real vehicle
model, and to determine blockage free aerodynamic
drag.

Numerical Methods
Governing equations and discretization:
The governing equations being solved in the LES are
spatially filtered and Navier-Stokes equations:
u i
0,
(1)
x i
u i

SGS S ij ,

ui u j
2
x i
x j
t
x j

(2)

The bar over the physical quantity indicates the spatial


filtering operation for LES. The filtered strain rate
tensor S ij and pressure P in Eq. (2) are expressed as

S ij

1 u j u i

2 xi x j

(3)

P p / uiu j ui u j / 3 ,

(4)

In Eq. (2), the last term on the right represents the


effect of sub-grid-scale (SGS) turbulence, which was
modeled under the eddy viscosity assumption. The
conventional Smagorinsky model (Smagorinsky. 1963)
was used, and the eddy viscosity coefficient was
modeled as

SGS C s f d 2 2 S ij S ij ,

(5)

where is the volume of numerical element, and


model coefficient C s is set to be 0.15, which is
generally suitable for external flows. The damping of
the turbulent effect near a wall boundary is explained
by the Van-Driest type damping function as follows:
f d 1 exp

l
,
25

(6)

where l is the distance from the wall in wall


coordinates.
The governing equations were discretized by using the
vertex-centered unstructured finite volume method. The
second-order central differencing scheme was applied
for the spatial derivatives and blending of 5% firstorder upwind scheme for the convection term was
employed for numerical stability. The third-order
upwind scheme was adopted for the spatial derivative
far away from vehicle, where coarser grid was
allocated. For time marching, the third-order AdamsMoulton semi-implicit scheme was used. Pressurevelocity coupling was preserved by using the
Simplified Marker and Cell (SMAC) algorithm.

Fig 2. Space elements on the vehicle components


The triangle element on the surface of the vehicle is
indicated in Fig 2. The fluid space was decomposed by
tetrahedral elements. To maintain finer resolution
around the vehicle, hierarchical allocation is carried
out, as shown in Fig 3.

Target Vehicle Model


The configuration of the two different full-scale heavyduty trucks is shown in Fig 1.The surface of the vehicle
is reproduced by about 1.5 million triangle meshes. To
reproduce the fine structure, the surface resolution is
around 5 to 10 mm around the side mirror, and
relatively fine elements are allocated around the cabin.
The engine and power train is reproduced by the
moderate elements with the resolution of 20 to 50 mm.
Larger elements are allocated to reproduce the cargo
panel.

Fig 1. Models with (Left) and without (Right) the air


deflector

Fig 3. Space elements around the vehicle

Computational Domains and Boundary


Conditions
To validate our numerical method based on the DNWLLF wind tunnel data, we adopted the same wind
tunnel geometry in the simulation, as shown in Fig 4(a),
in which the blockage ratio of the projected frontal
area of the vehicle and the cross sectional area of the
test section is about 10%. To compare the corrected
drag with the ideally obtained value, we have also
conducted the case with the blockage ratio of around
1%, as shown in Fig 4(b). In both cases, a uniform
velocity distribution U 0 is defined at the inlet (about 22
m/s and 25 m/s in the DNW and ideal cases,
respectively) about 40 m upstream of the vehicle. All
velocity components were gradient-free for the
steamwise direction at the outlet. The log-law profile
was assumed on the velocity and surface friction on the
wall was estimated and directly imposed as Neumann
boundary condition. In DNW case, shown in Fig 4(a),
the exit of the wind tunnel nozzle is located 3.79 m
windward of the vehicle and diffuser type collected is
mounted after the test section.

(a)

Mercker correction:
Mercker [3] provides a blockage correction equation
based on the work of Lock for solid blockage and on
the work of Maskell, Thom, and Glauert for wake
blockage. There is important thing of Mercker
correction is considered about yaw angle effect, vehicle
geometry as well as wind tunnel geometry. The
correction expressed by the ratio of dynamic pressure is

(b)

2 AM .2VM
qc
2A
1 K3
M
q
2 AN
3/ 2
L p .2VM .(2 AN )

2 AM
1
. CDm (0)
2 AN
4

(9)
where s is solid blockage factor, w is wake blockage
factor, AM is vehicle frontal area at

deg. yaw angle,

VM is vehicle volume, and CDm (0) is the measured


drag coefficient at 0 deg. yaw angle.
Mercker recently modified his procedure in order to
consider the static pressure change over the test section
at empty state. The drag coefficient is fully corrected in
order to consider wake distortion and horizontal
pressure buoyancy by
C D T C D uw C pw C DHB ,
(10)

Fig 4. Computational domains: DNW geometry,


(b) Ideal

Review of Wind-Tunnel Blockage


Correction Equations
It is defined that the total blockage correction factor is
the sum of velocity acceleration (blockage factor)
caused by solid and wake blockage; however these are
more difficult factors to assess for unusual geometries
such as the heavy-duty truck with including
complicated detail characteristics features and the
associated flow fields around them.

Maskell correction:

(7)

C D w C D uw / qc / q ,
where

qc is

(11)

Wake distortion term:


C pw C pwc C pmb ,

(12)

where C pwc is the pressure coefficient in empty wind


tunnel at the location of wake closure and C pmb is the
pressure coefficient in empty wind tunnel at the
location of the base of the model.
Horizontal pressure buoyancy term:

CDHB 1.75 / AM . VM / 2.G ,

Maskell [2] was the first to address the problems with


non-streamline flow bodies, such as bluff-body testing
in closed-wall wind tunnel and that partially stalled
shapes such as wings. Maskells theory holds for close
axis symmetric wake in three dimensional flows. The
correction of the dynamic pressure ratio is as follows:

A
qc 5
1 C D uw M
q 2
AN

C D w C DT / qc / q ,

corrected

(8)
dynamic

pressure,

q is

uncorrected dynamic pressure, AM the flat plate area,

AN the wind tunnel working section cross sectional


area, CDw is corrected drag coefficient and CDuw is
uncorrected wind-axis drag coefficient .

(13)

Evaluation of the Correction Equation


The time-averaged aerodynamic drag coefficient shown
in Table 1 was obtained by DNW wind tunnel
geometry and ideal condition wind tunnel simulation
for the model with air deflector. All the aerodynamic
coefficients are normalized for confidentiality reasons.
In Table 2, the aerodynamic drag is corrected by
Maskell and Mercker methods. The corrected results
were obtained by using drag coefficient from DNW
geometry simulation. The corrected value by Mercker
method shows good agreement with the ideal value,
when consider solid blockage, wake blockage and
horizontal buoyancy term. Also Maskells correction
equation has a similar trend in 0 degree yaw angle and
it shows over correction in 10 degree yaw angle. On the
other hand, when taken into account all of the effects

including wake distortion term in the Merckers


method, drag is overcorrected.
Table 1: Normalized drag coefficient for the model with air
deflector (0 and 10 deg. yaw angle)

Normalized drag coefficient (CFD)


Yaw angle(deg.) DNW geometry Ideal LES
0
1.000
0.801
10
1.716
1.297
Table 2: Normalized corrected drag coefficient for the model
with air deflector by correction Eq. (0 and 10 deg. yaw angle)

Normalized Corrected Drag coefficient based on


Mercker Equation
Factors considered
0 deg.
10 deg.
Solid + Wake blockage
0.847
1.329
Solid + Wake blockage + HPB 0.789
1.278
Solid + Wake blockage + HPB 0.693
1.189
+ Wake distortion

relevant length behind the vehicle over which the


pressure gradient effect has to be calculated. This
location is measured up to a first order of
approximation, which occurs where C PT 0 , here

CPT is total static pressure coefficient. Recently,


Mercker [4] has noticed that the true length (sensitivity
length) can be determined exactly by generating a
second gradient in the empty test section. It is turned
out that this length is usually shorter than the distance
between vehicle base location and wake closure
location. For that reason, when using the wake closure
location, the buoyancy effect on the wake
overcorrected the measured drag coefficient. We can
say that at this moment we are yet in the position to
determine the sensitivity length on a pure theoretical
base other than on experimental grounds.

Conclusions
Normalized corrected Drag coefficient based on
Maskell Equation
Factor considered
0 deg.
10 deg.
Solid blockage
0.787
1.174
Table 3 and Table 4 present the result obtained with
model without the air deflector at 0 degree yaw. We
can see that aerodynamic drag corrected by Merckers
method shows a better agreement when excluding the
wake distortion effect. The corrected Maskells method
resulted in an over correction. From our CFD results,
we realize that Maskells method shows the worst
result in non-zero yawing.
Table 3: Normalized drag coefficient for the model without
air deflector (0 deg. yaw angle)

Normalized drag coefficient (CFD)


Yaw angle(deg.) DNW geometry Ideal LES
0
1
0.812
Table 4: Normalized corrected drag coefficient for the model
without air deflector by correction Eq. (0 deg. yaw angle)

Mercker
Factors considered
Solid + Wake blockage
Solid + Wake blockage + HPB
Solid + Wake blockage + HPB
+ Wake distortion

0 deg.
0.848
0.802
0.762

Maskell
0 deg.
0.764

Concerning the relatively large discrepancies between


the ideal drag coefficient and the corrected drag
coefficient by Maskells method, one possible
explanation is that the Maskells correction equation is
based on
simplified geometry (symmetric wake
structure at rear section of a circular flat plate) and do
not account for the effect of yawing. To consider the
wake distortion in Merckers method, which depends
on approximation location of the wake closure, the

Computational fluid dynamics (CFD) flow simulations


have successfully been performed using large eddy
simulation to assess the wind tunnel blockage effects
on real vehicle model. The Maskells correction
equation for aerodynamic drag is working well for the
model with the air deflector without considering the
yawing effect. The drag coefficients under the effect of
yaw angle obtained with the correction method
proposed recently by Mercker had shown good
agreement with the ideally obtained values in both
models. But further investigation is needed for wake
induced drag increment for steep pressure gradient in
Merckers procedure.

Acknowledge
This work was supported by the Industrial Technology
Research Grant Program in 2007 from the New Energy
and Industrial Technology Development Organization
(NEDO) of Japan. This study was conducted in a
collaborative research project with Isuzu Advanced
Engineering Center Ltd., and the geometry data and
experimental data received are greatly acknowledged.

References
1.

K.R. Cooper et al.: "Closed-Test-Section Wind Tunnel


Blockage Corrections for Road Vehicles."SAE
International. SP-1176.
2. Maskell, E.C. A Theory of the Blockage Effects on
Bluff Bodies and Stalled Wings in a Closed Wind
Tunnel. ARC R&M 3400, Nov. 1963.
3. E. Mercker et al.: The Influence of a Horizontal
Pressure Distribution on Aerodynamic Drag in Open
and Closed Wind Tunnels, SAE 2005-01-0867, 2005.
4. E. Mercker, BMW AG, Germany and Dr. Koro Kitoh,
Kozo Kitoh Technology. Inc, 2011: Private
communication.

You might also like