You are on page 1of 9

Licensed to Missouri University of Science and Technology

Licensed from the SAE Digital Library Copyright 2009 SAE International
E-mailing, copying and internet posting are prohibited
Downloaded Sunday, October 25, 2009 6:06:31 PM

SAE TECHNICAL
PAPER SERIES

983041

Viscous-Flow Simulation of an Open-Wheel


Race Car
Joseph Katz
SDSU

Hong Luo, Eric Mestreau and Joseph Baum


SAIC

Reinald Lhner
George Mason Univ.

Reprinted From: 1998 Motorsports Engineering Conference Proceedings


Volume 1: Vehicle Design and Safety
(P-340/1)

Motorsports Engineering
Conference and Exposition
Dearborn, Michigan
November 16-19, 1998
400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A.

Tel: (724) 776-4841 Fax: (724) 776-5760

Author:Gilligan-SID:13235-GUID:24526534-131.151.244.7

Licensed to Missouri University of Science and Technology


Licensed from the SAE Digital Library Copyright 2009 SAE International
E-mailing, copying and internet posting are prohibited
Downloaded Sunday, October 25, 2009 6:06:31 PM

The appearance of this ISSN code at the bottom of this page indicates SAEs consent that copies of the
paper may be made for personal or internal use of specific clients. This consent is given on the condition,
however, that the copier pay a $7.00 per article copy fee through the Copyright Clearance Center, Inc.
Operations Center, 222 Rosewood Drive, Danvers, MA 01923 for copying beyond that permitted by Sections 107 or 108 of the U.S. Copyright Law. This consent does not extend to other kinds of copying such as
copying for general distribution, for advertising or promotional purposes, for creating new collective works,
or for resale.
SAE routinely stocks printed papers for a period of three years following date of publication. Direct your
orders to SAE Customer Sales and Satisfaction Department.
Quantity reprint rates can be obtained from the Customer Sales and Satisfaction Department.
To request permission to reprint a technical paper or permission to use copyrighted SAE publications in
other works, contact the SAE Publications Group.

All SAE papers, standards, and selected


books are abstracted and indexed in the
Global Mobility Database

No part of this publication may be reproduced in any form, in an electronic retrieval system or otherwise, without the prior written
permission of the publisher.
ISSN 0148-7191
Copyright 1998 Society of Automotive Engineers, Inc.
Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE. The author is solely
responsible for the content of the paper. A process is available by which discussions will be printed with the paper if it is published in
SAE Transactions. For permission to publish this paper in full or in part, contact the SAE Publications Group.
Persons wishing to submit papers to be considered for presentation or publication through SAE should send the manuscript or a 300
word abstract of a proposed manuscript to: Secretary, Engineering Meetings Board, SAE.

Printed in USA

Author:Gilligan-SID:13235-GUID:24526534-131.151.244.7

Licensed to Missouri University of Science and Technology


Licensed from the SAE Digital Library Copyright 2009 SAE International
E-mailing, copying and internet posting are prohibited
Downloaded Sunday, October 25, 2009 6:06:31 PM

983041

Viscous-Flow Simulation of an Open-Wheel Race Car


Joseph Katz
SDSU

Hong Luo, Eric Mestreau and Joseph Baum


SAIC

Reinald Lhner
George Mason Univ.
Copyright 1998 Society of Automotive Engineers, Inc.

ABSTRACT

initially inviscid methods were used, with some success


[4], for various race cars and road vehicles with large
regions of attached flows. Because of the massive flow
separations, particularly those created by the large
wheels, viscous flow simulation is necessary for open
wheel race cars. Consequently, the solution of the full
Navier-Stokes (N.S.) equations is required for the flowfield surrounding the complex geometry of an INDY-type
race car, such as the one shown in Fig. 1.

A numerical solution based on the Navier-Stokes equation, combined with unstructured grid mesh, was used to
model an open wheel race car. The solution is based on
a fast, matrix-free, implicit method, with relatively low
storage requirements, resulting in solution times up to an
order of magnitude smaller than other numerical solutions. The computations provide details on the flow field
around the car and a complete pressure distribution on
the vehicles surface. The calculated results may be used
as a supplementary tool for wind tunnel or road testing
and can provide information, such as the underbody flow,
which is difficult to evaluate experimentally. One of the
primary advantages of such a viscous flow simulation is
the ability to model wheel rotation and to detect regions
of flow separation, particularly on the suction side of the
front and rear wings.

INTRODUCTION
The complex nature of the flowfield over open-wheel race
car configuration often results in numerous non-linear
effects which complicate the estimation of expected performance-gains due to proposed design modifications.
Traditionally, both wind-tunnel and road tests are used
extensively during the aerodynamic development of such
a vehicle. Those methods have matured to provide
dependable integral results such as the lift and drag but
quite often cannot provide sufficient details on the flowfield to explain why certain trends are observed. Therefore, a design tool that can provide detailed flowinformation on and off the surface of the vehicle, can help
explain some of those nonlinear effects and shorten the
design cycle. To fill this void, recently, the use of complementary computational methods has increased [1-3] and
their significance as an additional independent design
tool has been gradually established. Earlier computational tools were limited by computer performance and

Figure 1. Solid body model of a generic INDY race car.


The objective of the present work is to demonstrate the
applicability of the approach as a complementary tool,
particularly in areas where the experimental approach
has limited capabilities. For example, flow-visualization
inside the diffusers under the car, as well as tracing the
front wing wake in a high-Reynolds number flow, is close
to impossible in a wind tunnel experiment. Similarly, evaluating the pressure distribution over a large number of
proposed wing configurations, in a small-scale test, may
become an endless chore. In all these cases, the use of

Author:Gilligan-SID:13235-GUID:24526534-131.151.244.7

Licensed to Missouri University of Science and Technology


Licensed from the SAE Digital Library Copyright 2009 SAE International
E-mailing, copying and internet posting are prohibited
Downloaded Sunday, October 25, 2009 6:06:31 PM

found in the 1996 rulebook [10] of the sanctioning organization (CART).

computational methods can easily fill the required information gap.

The grid generator used is FRGEN3D [9]. This unstructured grid generator is based on the advancing front
method. After defining the surface of the domain to be
gridded, the surfaces are triangulated. Thereafter, the
face that forms the smallest new element is deleted from
the front, and a new element is added. This process is
repeated recursively until no more faces are left in the
front. Results of the grid generation are presented in Fig.
2. and the triangular surface elements density can be
easily adjusted in areas where larger resolution is
required, without dramatically increasing the total number
of grid points. The mesh used in this computation
assumes a symmetrical model and contains 8,314,454
tetrahedral elements, 1,459,199 grid points, and uses
174,134 boundary points on the car surface and the far
field (per one half of the model).

METHODOLOGY
The present work is based on solving the finite element
equivalent of the N.S. equations with the flux-corrected
transport concept (Ref. 5). An unstructured mesh is used
which makes the grid generation around complex configuration much simpler than with other structured grid
meshes. The process usually begins with the preprocessor FECAD that prepares the CAD surface for the grid
generator, FRGEN3D. The grid generator then creates
the computational mesh, which is used by the solver,
FEFLOW97/8, that solves the finite-element equivalent of
the N.S. equations. A post processor (FEPOST3D) is
then used to graphically display and analyze the computational results. These utilities are briefly described in the
following paragraphs.
Generation of the vehicle surface model is still one of the
largest tasks in the process of obtaining a numerical solution. The experience gained over several years in the
generation of surface and volume meshes for the simulation of flows about such complex geometric structures as
airplanes, tanks, trains, cars and trucks [6-8], has shown
that once the surface and volume mesh generation has
been automated, the surface description (i.e., point, line
and patch definition), as well as the correct definition of
boundary conditions, become the dominant time consuming task. This information led to the development of
the preprocessor FECAD, a suite of efficient, userfriendly utilities that allows the quick production of
FRGEN3D-compatible, error-free input. In addition to
basic CAD-CAM operations (shrinking, translating, rotating, surface lofting, etc.), FECAD also eases the merging
of several surface parts into one cohesive, well-defined
input-file. This allows the merging of files produced by different users and/or different surface generators. Most
importantly, FECAD allows graphical interrogation of the
surface data, and has many built-in diagnostics to avoid
undesirable features such as doubly defined points, isolated points or lines, badly defined lines or surfaces, and
lines or surfaces that are directed incorrectly. FECAD
also allows the specification and visualization of boundary conditions, saving the user many error-prone hours
during the later stages of a run. FECAD has proven to be
an invaluable aid when trying to construct an error-free
FRGEN3D-compatible input file in a matter of days or
even hours. In order to specify the desired element size
and shape distribution in space, a combination of background grids [9] and sources is employed. This task is
again performed within the point-and-click environment of
FECAD.

Figure 2. Unstructured grid around the open-wheel race


car model. To simulate a moving ground, the
wheels rotate in coordination with the speed of
the floor grid.
The flow solver employed was FEFLO97, a new 3-D ALE
edge-based [11] hydro-solver based on the Finite-Element Method Flux-Corrected Transport (FEM-FCT) concept [5], which solves the Euler and Reynolds-averaged
turbulent, Navier-Stokes equations. H-refinement is the
preferred approach for grid adaptation [12]. The highorder scheme used is the consistent-mass Taylor-Galerkin algorithm. Combined with a modified second-order
Lapidus artificial viscosity scheme, the resulting scheme
is second-order accurate in space, and fourth-order
accurate in phase. The spatio-temporal adaptation is
based on local H-refinement, where the refinement/deletion criterion is a modified H2-seminorm [12]. The critical
parameter for the refinement/deletion criteria is density.
Various turbulent models can be used, and the
model was tested successfully with this approach. For
the calculations presented here a full-size vehicle at 100
mph was simulated; however, no turbulence model was

The solid body model of a generic INDY car, created by


the FECAD preprocessor is shown in Fig. 1. The model
includes simulation of the internal flow across the radiator
ducts and the proper boundary conditions to simulate
wheel rotation and moving ground. More details on the
geometry and specific dimensions of this vehicle can be

Author:Gilligan-SID:13235-GUID:24526534-131.151.244.7

Licensed to Missouri University of Science and Technology


Licensed from the SAE Digital Library Copyright 2009 SAE International
E-mailing, copying and internet posting are prohibited
Downloaded Sunday, October 25, 2009 6:06:31 PM

Typical streamline traces in the flowfield nearby the car


model are presented in the following six figures. These
figures, as viewed in the post-processor, are far more
detailed and use colors to identify the local velocity on the
streamline. Unfortunately, most of the resolution and the
aerodynamic information is lost while converting those
flow visualizations into the gray-scale images in this article. An example of such streamline traces is shown in
Fig. 3. The tracing begins behind the front wheels and
the fluid particles are partially captured by the large separation zone behind the wheel. Some slow moving particles may rise in the flow behind the wheel and join the
high speed stream above it. To simulate actual road condition the tire had a flat contact patch with the road which
was further modified by two small wedges inserted
underneath. This was done to improve local numerical
stability and was plotted by the postprocessor as an elevated block (see Figs 3 and 4). Also, the effect of the
small turning vane in front of the side pod is clearly visible
I this figure. The flow is kept attached to the bodywork
and is less affected by the wheel-base separation. The
streamlines above the wheel seem to head inward and
eventually will end up flowing beneath the rear wing.

used and the data represents a virtual high-Reynolds


number laminar flow case. When compared with experimental data, this temporary simplification typically results
in earlier flow separation on highly curved surfaces.
The FEFLO98 solver has recently been improved with a
new, fast, matrix-free implicit algorithm. This algorithm
solves an approximate system of linear equations which
arise from the Newton linearization by the GMRES (Generalized Minimum RESidual) algorithm with a LU-SGS
(Lower-Upper Symmetric Gauss-Seidel) preconditioner.
This preconditioner has been further modified to allow
the accurate solution of virtually incompressible flows
(i.e., down to Mach numbers of 0.00001). The most
remarkable feature of the new GMRES+LU-SGS method
is that the storage of the Jacobian matrix can be completely eliminated by approximating the Jacobian with
numerical fluxes, resulting in a matrix-free implicit
method. Numerical tests demonstrate a factor of 10-15
increase in performance over the best current implicit
methods, without increasing memory requirements. In
this particular case, the residual is reduced by three
orders of magnitude within 200 time steps and the whole
computation requires about 5 hours of CPU time (on a
Cray C-70 computer).
Post-processing was performed with the FEPOST3D and
FEMOVIE packages [13]. FEPOST3D performs all CPUintensive filtering operations on the Supercomputer. Only
the plane or surface data are sent back to the graphics
workstation for plotting.

Figure 4. Streamline traces released in the vehicle


symmetry plane.
The streamline traces shown in Fig. 4 provide a general
indication about the flow direction in the symmetry plane.
The raised nose design clearly allows sufficient flow
under the vehicle, but most of the flow moving above the
nose will end up near the drivers helmet. The rear wing
clearly changes the streamline curvature, indicating the
large level of downforce created. However, the streamlines leaving the trailing edge flap are not parallel to the
flap due to local flow separation there. This separation
may be exaggerated by the current computation but
some partial flow separation was observed during road
testing with the higher downforce settings. The lack of
streamlines behind the lower portion of the vehicle may
indicate the high bluff body drag created.

Figure 3. Streamline traces in the flow under the car.

RESULTS
Results of the computations are divided into two sections.
The first is the off-body flow visualization and the second is the surface pressure distribution. Flow visualizations are widely used during wind-tunnels experiments;
however, in some areas conventional methods are not
always effective. Similarly, surface pressure measurements may be measured in a wind tunnel but are not
widely used in race-car wind tunnels because of the short
development time requirement.

The information about the orientation of the flow ahead of


the vehicle is relevant to both downforce and cooling considerations. Fig. 5 and Fig. 6 demonstrate how some of
this information can be obtained by using numerical flow

Author:Gilligan-SID:13235-GUID:24526534-131.151.244.7

Licensed to Missouri University of Science and Technology


Licensed from the SAE Digital Library Copyright 2009 SAE International
E-mailing, copying and internet posting are prohibited
Downloaded Sunday, October 25, 2009 6:06:31 PM

above the rear bodywork and travels under the large rear
wing. The effect of the small flow turning device above
the rear wing may be clarified by observing the streamlines in this area. Because of the locally horizontal flow it
has the potential to create some downforce, and at the
same time reduce the rear wheel drag by diverting the
flow away from it.

visualizations. For example, the raised central portion of


the front wing allows more flow to move under the car,
which eventually contributes to more airspeed there and
more downforce. But, more importantly, this portion of the
flow seems to reach the cooling inlets, directly affecting
the side pod design. The streamlines in Fig. 6 show that
some of the flow at the wing root still reaches the cooling
ducts, but most of the streamlines travel above the vehicle side pod. Once the airflow moves past the side pods
(Figs. 5 and 6), it is channeled below the rear wing. A
careful examination of the streamlines near the front wing
tip, in both figures, indicates that the tip-vortex stretches
inside the front wheels and then moves outside and
upward, missing the rear wing.

Figure 7. Streamline traces released outside of the front


wing side fin.
The top view of the flow approaching the rear wing is
shown in Fig. 8. The boat-tail shape of the body and the
large suction under the rear wing is forcing the flow to
move inward between the rear wheels and the body. This,
in effect, creates an angularity in the direction of the flow
reaching the rear wheel and the rear separation bubble
faces inward. Further behind the car, after passing the
rear wing, the streamlines are closing inward quickly. This
is most likely a reaction to the low pressure base flow (or
bluff body effect) of the not so streamlined vehicle and
its wheels and due to the flow induced by the rear wing
vortices.

Figure 5. Streamline traces released at the front wing


leading edge level.

In the second part of this article, calculated surface pressure data is presented. The surface pressure distribution
is a direct outcome of the copmputations and readily
available on all surfaces. By integration, it can be used to
determine aerodynamic loads on various body panels
and can be used to redesign a particular surface shape
(e.g. to reduce flow separation). As an example, the pressure distribution along the ground-plane symmetry line is
presented in Fig. 9. The front suction peak is created by
the front wing, while the second one is a result of the flow
accelerating between the front wheels. The largest suction peak is at the entrance to the underbody tunnels
(which still exist in INDY cars), and the pressure coefficient then gradually increases to values near zero behind
the vehicle. The experimental data shown by the (*) symbols was collected by a static pressure rake behind the
car, and it agrees reasonably well with the computed values.

Figure 6. Streamline traces released above the front


wing leading edge plane.
The high suction of the rear wing is evident in Fig. 7 as
well. Here the streamlines were released in a plane outside the front wheels. The flow initially is displaced outward by the large rotating wheels; however, the suction
behind the front wheels turns some of the streamlines
inward. A large portion of the streamlines eventually rises

Author:Gilligan-SID:13235-GUID:24526534-131.151.244.7

Licensed to Missouri University of Science and Technology


Licensed from the SAE Digital Library Copyright 2009 SAE International
E-mailing, copying and internet posting are prohibited
Downloaded Sunday, October 25, 2009 6:06:31 PM

ues above the wheel. The small dip near the top of the
tire is probably due to the flow separation line there. This
suction region inside the separated region, extends with
a fairly constant pressure coefficient behind the wheel,
clearly contributing to the drag force on this wheel.
The pressure distribution (or velocity distribution) on a
wing can serve to evaluate its performance and indicate if
its lift or drag can be altered according to a particular
design need. For example, the pressure distribution
along longitudinal sections of the race cars front and rear
wings is provided in the following three figures. The pressure distribution near the mid-line of the inner section
(with the shorter flap, see Fig. 1) is shown in Fig. 11.

Figure 8. Top view of the streamline traces approaching


the rear wing.

Figure 10. Polar description (with Cp=0 on the circular tire


outline) of the pressure coefficient along the
rear wheel centerline.

Figure 9. Pressure distribution on the moving ground


plane centerline.
Experimental evaluation of the pressure distribution on
moving surfaces, such as shown in Fig. 9, is quite difficult, particularly when it comes to rotating wheels. Fig. 10
demonstrates such results when using the computations
with a rotating surface. Here the pressure distribution
along the wheel centerline is presented in a polar diagram (note that the Cartesian ordinates serve to indicate
the Cp magnitude only, but the pressure distribution
shown is wrapped around the circular tire outline). In front
of the tire, at the lower part of the wheel, a high pressure
area exists (positive is inside the circle, on which Cp =0),
and the loop in the data shows the pressures on the
ground as well. As the flow accelerates over the wheel,
the pressure coefficient drops and reaches negative val-

Figure 11. Pressure distribution along an inner section


(with the shorter flap) of the front wing.

Author:Gilligan-SID:13235-GUID:24526534-131.151.244.7

Licensed to Missouri University of Science and Technology


Licensed from the SAE Digital Library Copyright 2009 SAE International
E-mailing, copying and internet posting are prohibited
Downloaded Sunday, October 25, 2009 6:06:31 PM

design for the front wing. The trailing edge flap angle for
this particular condition, seems to be set at a too high
value and the shown gap between the two elements is
probably too large. The interaction between the main
wing plane and its flap may be improved by closing the
gap somewhat.

The shape of the pressure distribution in Fig. 11 follows


current design trends with a large laminar portion on the
wings suction (lower) side. In this particular computation
(without the turbulence model) a laminar bubble seems to
form behind the leading edge, creating the wavy shape in
the Cp curve near the leading edge. The rear flap shows
similar trends and some visible trailing edge separation
as well. It is believed that in the high-speed (over 100
mph) and turbulent flow case, those laminar bubbles near
the leading edge will disappear.

The rear wing pressure distribution is presented in Fig.


13. In this case, the INDY-car regulations force the use of
a very long airfoil chord, resulting in a non-airplane like
pressure distribution. The suction peak at the front is typical to most airfoils at larger angles of attack; however,
the second suction peak at the aft section of the main
element is less traditional. This is a result of the sharp
turn of the streamlines near the gap region, causing a
local trailing edge separation. The flap, in this case, too,
is partially separated. This is standard practice in search
of maximum rear downforce at the cost of increasing
rear-wing drag. The streamline plot, presented in Fig. 14,
visualizes the flow in the symmetry plane of the rear
wing. The small trailing edge separation at the base of
the main wing element is hardly visible in this gray-scale
image, but the separation of the rear flap is more pronounced. These flow separations are amplified by the
absence of the turbulence model, but traces of it were
detected during full scale testing (by using tufts).

Figure 12. Pressure distribution along an outer section


(with the longer flap) of the front wing.

Figure 14. Streamlines released near the symmetry


plane of the rear wing.

CONCLUSIONS
Figure 13. Pressure distribution along the centerline of
the rear wing.

The numerical model used in this study demonstrates the


feasibility of solving the Navier-Stokes equations for a
complex race car configuration within practical time limits.
Computed results clearly demonstrate the ability to simulate viscous flow separation on the complex components
of open-wheel race cars (e.g the flow over rotating
wheels). When properly used, this approach can provide
the reasoning for planned model modifications. It can be
used to propose and numerically test numerous wing

A similar pressure distribution diagram, but along an


outer section (placed near the tip, at the center of the
larger flap) is shown in Fig. 12. In this case the flow on
the main element is attached, but partially separated on
the flap. The shape of the pressure distribution in these
two figures indicates a careful and well-balanced airfoil

Author:Gilligan-SID:13235-GUID:24526534-131.151.244.7

Licensed to Missouri University of Science and Technology


Licensed from the SAE Digital Library Copyright 2009 SAE International
E-mailing, copying and internet posting are prohibited
Downloaded Sunday, October 25, 2009 6:06:31 PM

shapes and to virtually investigate various ideas before


building a test model. Numerical results may be used in a
complementary manner with other experimental techniques because of the easily available pressure distribution and flow orientation information, particularly in areas
near the wings, under the car and near the rotating
wheels.

REFERENCES
1. Werner, F., Frik, S., and Schulze, J., Aerodynamic Optimization of the Opel Calibra ITC Racing Car Using Experiments and Computational Fluid Dynamics, SAE 98-0040,
Feb. 98, Detroit MI.
2. Axelsson, N., Ramnefors, M., and Gustafsson, R., Accuracy in Computational Aerodynamics Part 1: Stagnation
Presure, SAE 980037, Feb. 98, Detroit MI.
3. Perzon, S., Sjogren, T., and Jonson, A., Accuracy in Computational Aerodynamics Part 2: Base Pressure, SAE
980038, Feb. 98, Detroit MI.
4. Katz J., and Dykstra, L., "Application of Computational
Methods to the Aerodynamic Development of a Prototype
Race-Car," SAE Paper 942498, Proceedings of the 1994
Motor Sport Engineering Conf., P-287, pp. 161-169, , Dec.
5-8, 1994, Detroit, MI.
5. Lhner, R., Morgan, K., Peraire, J., and Vahdati, M., Finite
Element Flux-Corrected Transport (FEM-FCT) for the Euler
and Navier-Stokes Equations, Int. J. Num. Meth. Fluids 7,
1093-1109 (1987).
6. Baum, J.D. and Lhner, R., Numerical Simulation of Shock
Interaction with a Modern Main Battlefield Tank, AIAA-911666 (1991).
7. Baum, J.D., Luo, H., and Lhner, R., Numerical Simulation
of a Blast Inside a Boeing 747, AIAA--93-3091 (1993).
8. Baum, J.D., Luo, H., and Lhner, R., Numerical Simulation
of Blast in the World Trade Center, AIAA-95-0085 (1995).
9. Lhner, R., and Parikh, P., Three-Dimensional Grid Generation by the Advancing Front Method, Int. J. Num. Meth.
Fluids 8, 1135-1149 (1988).
10. C.A.R.T., Aerodynamic and Body Work Specifications,
Cart Rulebook 1996, Chapter 9.4.
11. Luo, H., Baum, J. D., and Lhner, R., Edge-Based Finite
Element Scheme for the Euler Equations, AIAA J. 32, 6,
1183-1190 (1994).
12. Lhner R., and Baum, J. D., Adaptive H-Refinement on 3D Unstructured Grids for Transient Problems, Int. J. Num.
Meth. Fluids 14, 1407-1419 (1992).
13. Lhner, R., Parikh, P., and Gumbert, C., Some Algorithmic
Problems of Plotting Codes for Unstructured Grids, AIAA89-1981-CP (1989).

Author:Gilligan-SID:13235-GUID:24526534-131.151.244.7

You might also like