Vous êtes sur la page 1sur 184

CALIFORNIA STATE UNIVERSITY NORTHRIDGE

Evaluation of Ground Effect on the Drag on an HPV Fairing Using CFD

A thesis submitted in partial fulfillment of the requirements


For the degree of Master of Science in Engineering,
Mechanical Engineering

By
Dimitry Tsybulevsky

May 2012

The Thesis of Dimitry Tsybulevsky is approved:

Susan Beatty, Eng.

Date

Mike Kabo, Ph.D.

Date

Robert G Ryan, Ph.D., Chair

Date

California State University, Northridge


ii

Acknowledgements
I would like to thank Dr. Robert Ryan for being my graduate advisor and
supporting me throughout this thesis. My thesis never would have been completed
without his help. I would also like to thank Professor Susan Beatty for helping me during
my time in California State University Northridge (CSUN) and being on my thesis
committee. Additionally, special thanks goes to Dr Mike Kabo for assisting me with the
application process for the graduate program in CSUN and being on my thesis
committee. Lastly, I would like to thank the Department of Mechanical Engineering at
CSUN for the encouragement and help to complete my Masters Degree in Mechanical
Engineering.

iii

Table of Contents
Signature page.ii
Acknowledgements....iii
List of Tables.................................vii
List of Figures...............................viii
Abstract.......................................xi
Chapter 1: Introduction1
1.1. Problem Statement..........1
1.2. Purpose of the Thesis.......1
1.3. Background Information..........3
1.3.1. Definition of Drag.....3
1.3.2. Definition of Ground Effect..................4
1.3.3. Definition of CFD and CFD History.....5
1.3.4. Drag Measurement Techniques Using CFD Approach.................6
1.3.5. Theoretical Values of Drag on the Ellipsoid body .......7
1.3.6. Drag Values on Variation With Ground Clearance....................12
1.4. HPV Fairing Geometry Description..... ....15
1.5. Organization of the Thesis.....16
Chapter 2: Importation of Solid Model into ANSYS and Mesh Definition..18
2.1. Meshing and Preprocessing...............18
2.2. Modeling of the HPV Fairing and Ellipsoid Geometries in SolidWorks..20
2.3. Importing the Model into ANSYS WORKBENCH from SolidWorks.22
2.3.1. Extracting A Fluid Volume for the Models.....24

iv

2.3.2. Opening the Models in ANSYS ICEM CFD..31


2.3.3. Preparing the Geometry for Meshing......32
2.3.4. Generating the Initial Mesh Using Octree Mesh Approach and Applying
the Correct Mesh Size.........34
2.3.5. Generating the Tetra/Prism Mesh Using Delaunay Mesh Approach......40
2.3.6. Smoothing the Mesh to Improve Quality........41
2.4. Exporting the Mesh into ANSYS FLUENT..........43
Chapter 3: FLUENT Setup and Application of Spalart-Allmaras Turbulence Model..47
3.1. Background Information in Computational Software and Methodology..47
3.2. Turbulence Model..48
3.2.1. Spalart Allmaras Turbulence Model...........49
3.3. Application of FLUENT Setup......55
3.3.1. Initial Setup.........55
3.3.2. Boundary Condition........59
3.3.3. Solution Setup and Mesh Adaption.................64
3.4. Solution to the Problem............70
3.4.1. Graphical and Numerical solutions.........71
3.5. Drag Calculation....71
Chapter 4: Baseline Solution and Calibration of FLUENT.......75
4.1. FLUENT Calibration Using Flat Plate......76
4.2. FLUENT Calibration Using Oblate Ellipsoids..80
4.2.1. Results for Oblate Ellipsoids.......83
4.3. Comparison between Hoerners data and CFD data......88

Chapter 5: HPV Fairing Results....93


5.1. HPV CFD Test Results......93
5.1.1. HPV Fairing Benchmark Results .......94
5.1.2. HPV Fairing at Different Ground Proximities Results.................106
5.1.3. Ground Clearance Effect on Pressure and Skin Frication.....................113
5.1.4. Ground Clearance Effect on Drag and Lift .....119
5.2. Estimation of Discretization Error...124
5.2.1. Discretization Error Calculation........126
5.3. Tradeoff Study Between Ground Clearances Drag and Stability for a Typical
HPV.....128
Chapter 6: Conclusion..133
References........135
Appendix A......138
Appendix B......148
Appendix C......163

vi

List of Tables
Table 2-1....30
Table 3-1....60
Table 3-2....61
Table 3-3....62
Table 3-4....65
Table 3-5....66
Table 4-1....78
Table 4-2....79
Table 4-3....89
Table 4-4....89
Table 4-5....89
Table 4-6....90
Table 4-7....92
Table 5-1....97
Table 5-2....98
Table 5-3......108
Table 5-4......124
Table 5-5......124
Table 5-6......128
Table 5-7......132
Table C-1.....163
Table C-2.....164
Table C-3.........165
Table C-4.....166
Table C-5.....167
Table C6......168
Table C7......169
Table C-8.....170
Table C-9.....171
Table C-10...172

vii

List of Figures
Figure 1-1.4
Figure 1-2.9
Figure 1-3...10
Figure 1-4...12
Figure 1-5...14
Figure 1-6...15
Figure 2-1...19
Figure 2-2...20
Figure 2-3...21
Figure 2-4...23
Figure 2-5...24
Figure 2-6...25
Figure 2-7...26
Figure 2-8...26
Figure 2-9...27
Figure 2-10.....28
Figure 2-11.....29
Figure 2-12.....30
Figure 2-13.....31
Figure 2-14.....32
Figure 2-15.....32
Figure 2-16.....35
Figure 2-17.....36
Figure 2-18.....36
Figure 2-19.....38
Figure 2-20.....39
Figure 2-21.....41
Figure 2-22.....42
Figure 2-23.....43
Figure 2-24.....44
Figure 2-25.....45
Figure 2-26.....45
Figure 2-27.....46
Figure 3-1...56
Figure 3-2...57
Figure 3-3...58
Figure 3-4...59
Figure 3-5...63
Figure 3-6...64
Figure 3-7...67
Figure 3-8...69
Figure 3-9...69
Figure 3-10.....70
Figure 3-11.....72
Figure 4-1...77
viii

Figure 4-2...79
Figure 4-3...80
Figure 4-4...82
Figure 4-5...84
Figure 4-6...85
Figure 4-7...86
Figure 4-8...87
Figure 4-9...88
Figure 4-10.....91
Figure 5-1...95
Figure 5-2...96
Figure 5-3...96
Figure 5-4...97
Figure 5-5...98
Figure 5-6...99
Figure 5-7.....100
Figure 5-8.....101
Figure 5-9.....101
Figure 5-10...102
Figure 5-11.......103
Figure 5-12...103
Figure 5-13...104
Figure 5-14...105
Figure 5-15...106
Figure 5-16...107
Figure 5-17...108
Figure 5-18...109
Figure 5-19...110
Figure 5-20...................111
Figure 5-21...................112
Figure 5-22...................112
Figure 5-23...................114
Figure 5-24...................115
Figure 5-25...................116
Figure 5-26...................117
Figure 5-27...................118
Figure 5-28...................119
Figure 5-29...................120
Figure 5-30...................122
Figure 5-31...................123
Figure 5-32...................130
Figure 5-33...................131
Figure 5-34...................132
Figure A-1....................122
Figure A-2....................138
Figure A-3....................141
ix

Figure A-4....................144
Figure B-1....................148
Figure B-2....................148
Figure B-3....................149
Figure B-4....................149
Figure B-5....................150
Figure B-6....................150
Figure B-7....................151
Figure B-8....................151
Figure B-9....................152
Figure B-10..................152
Figure B-11..................153
Figure B-12..................153
Figure B-13..................154
Figure B-14..................154
Figure B-15..................155
Figure B-16..................155
Figure B-17..................156
Figure B-18..................156
Figure B-19..................157
Figure B-20..................157
Figure B-21..................158
Figure B-22..................158
Figure B-23..................159
Figure B-24..................159
Figure B-25..................160
Figure B-26..................160
Figure B-27..................161
Figure B-28..................161
Figure B-29..................162
Figure B-30..................162

Abstract
Evaluation of Ground Effect on the Drag on an HPV Fairing Using CFD
By
Dimitry Tsybulevsky
Master of Science in Mechanical Engineering

The purpose of this study was to evaluate the ground effect on the Human
Powered Vehicle (HPV) Fairing with different ground clearances, and its effect on drag
using Computational Fluids Dynamics (CFD) software. The short term goal of this thesis
was to use the CFD software package ANSYS FLUENT, to find how the ground
clearance of the 2010 version of the HPV fairing affects the overall drag and to an
optimal ground clearance for the vehicle. The long term goal was to create a guide to help
future students use ANSYS FLUENT and other ANSYS software to create mesh and
CFD studies to find external forces such as drag and lift coefficients on objects moving
through a fluid.
In order to create a good computational mesh for the HPV fairing flow field, the
mesh was first created for standard geometries, i.e. flat plate and oblate ellipsoids. Drag
values computed for various meshes were compared to known drag values for those
geometries. The results for the flat plate matched within 3.5% of the theoretical results,
and for the oblate ellipsoids the difference was less than 5.6% from experimental values.
This process helped to optimize the final mesh settings for the HPV fairing and find
acceptable results for the drag coefficient with the fairing at different ground clearances.

xi

As mentioned previously, a long term goal for this thesis was to create a tutorial
on how to use ANSYS and FLUENT to create good CFD studies. The tutorial can be
used with future California State University, Northridge (CSUN) senior design teams to
create body geometries and effectively to accurate results for drag and lift on various
bodies. This tutorial can also help with regard to importing the geometry from CAD
software and performing the correct model setup in ANSYS.
The study for the HPV was conducted as a function of h/L, where h is the ground
clearance and L is the length of the HPV fairing. (L= 99 inches and was constant). The
ground clearance ranged from 3 to 18 inches including two baseline tests, at 30 and 297
inches away from the ground. All of the results are provided in terms of the streamlines,
pressure and velocity magnitude fields, and vorticity contours.
The goal was to see how high the body had to be off the ground to eliminate the
drag ground effect. It was found that the fairing had to be at least 18 inches of from
ground in order to see a significant reduction in ground effect. Additionally a trade off
analysis was conducted on the HPV fairing to balance the speed benefit from high ground
clearance with vehicle stability during cornering. However, the height required to
minimize the ground effect was impractical for the HPV competition due the Center of
Gravity (CG) considerations.

xii

Chapter 1: Introduction
1.1.

Problem statement
The aerodynamics of human powered vehicles (HPVs) is greatly influenced by

the shape of the body and the proximity of the ground to the surface of the HPV
bodywork. In most cases the airflow between the ground and HPV bodywork results in a
drag increase known as the ground effect. Approaches to lessen this effect fall into two
categories: a) creating a specialized fairing skirt which helps to direct the airflow away
from the underside of the vehicle; or b) increasing the height of the vehicle from the
ground. Neither of those strategies is perfect; each strategy has its upside and its
downside with respect to vehicle performance.
1.2.

Scope of the Thesis


The main goal of this thesis is to conduct a computational fluid dynamics (CFD)

study on an HPV fairing by using ANSYS 12.1 and FLUENT in the Mechanical
Engineering Design center at California State University Northridge (CSUN). This study
analyzes airflow around a typical HPV fairing geometry and assesses the impact of the
ground effect at typical HPV speeds. In addition, this study is designed to use the oblate
ellipsoid and the flat plate as a calibration tool for the HPVs fairing mesh, boundary
conditions and FLUENT setup. Then the experimental results found in Fluid Dynamics
of Drag by Hoerner

[11]

are compared to the CFD results from FLUENT for the oblate

ellipsoid to make sure that the software computationally precise.


To accomplish these objectives the SolidWorks model created by 2010 CSUNs
HPV design team was imported into ANSYS 12.1 and modified to be used within
ANSYS-FLUENT. The geometry was cleaned within ANSYS 12.1 WORKBENCH
1

Geometry Design-Modeler; then the model was imported into ANSYS ICEM to create
the mesh that was used by FLUENT. The mesh incorporates an estimation of boundary
layer thickness to insure that sufficient points were used near the HPVs fairing surface to
accurately predict velocity gradients in this region.
Initially a study was performed on an ellipsoid geometry, which is somewhat
similar to the shape of an HPV, and for which published drag data is available. In
addition, velocities were chosen to match Reynolds numbers with available data. Using
the ellipsoid geometry, a strategy was developed to optimize the program settings to get
an effective convergence and solution accuracy in terms of drag force. This included
running inviscid flow cases, using coarser mesh for the preliminary calculations, and then
using FLUENT mesh refinement capabilities. In addition, different turbulence models
such as the Spalart-Allmaras turbulence (SA) model and k- model within FLUENT were
tried to assess the turbulence models effect on solution convergence and drag
calculations. This study was conducted using several different flow conditions and mesh
configurations to determine their effect on the calculated drag values.
The analysis was conducted on the 2010 HPV geometry at several different flow
velocities with a maximum flow velocity of approximately 40 mph (58.67

). These

speeds corresponded to a Reynolds number range of approximately 5 105 to 3 106 .

That means the majority of the flow over the HPV fairing after the expected boundary
layer transition point was in the turbulent region.
Finally, a study was conducted to assess the impact of geometry changes on
computed drag, i.e. changing the proximity of the HPV fairing to the ground surface.

Analyses were run for ground clearance of 3, 6, 9, 12, 15, 18, 30 inches and a freestream
case of 297 inches above the ground.

1.3.

Background Information

1.3.1. Definition of Drag


Drag refers to the forces that oppose the relative motion of an object through a
fluid, either gas or liquid. Drag forces only act in the direction opposite to velocities not
the oncoming flow velocity (or upstream velocity U). For a 3-D object moving through a
fluid, the drag is the sum of forces due to pressure differences in the flow field (pressure
drag) and shear forces on the objects surface (friction drag).
Drag force has been found to be dependent on a fluids density (), object area
(A), flow velocity (U) and a dimensionless drag coefficient (C D ), expressed by the
following drag equation:
1

= 2
2

(1-1)

The drag coefficient is a function of object shape and Reynolds number, and is
usually determined experimentally or by CFD analysis. The area can either be the surface
or wetted area, or the projected frontal area depending on the source of the drag
coefficient values. Generally the wetted area is used if the total drag is dominated by
friction drag.

Figure 1-1 a shows basic example of drag generated by a solid body moving
through a fluid.

Figure 1-1:
Example of drag generated by solid object
(Adapted from http://www.grc.nasa.gov/WWW/K-12/airplane/drag1.html )

1.3.2. Definition of Ground Effect


Ground effect is a term applied to a series of aerodynamic effects that are
important in the automotive and aerospace industries. These effects usually cause an
increase in drag force and a decrease in lift force (i.e. increase down force). Ground
effects relevant to the automotive industry are due to the proximity of the underside of
the moving vehicle to the stationary road surface. The ground effect is easily visualized
by taking a canvas tarp out on a windy day and holding it close to the ground; when the
canvas gets close enough to the ground it will suddenly be sucked downward due to the
lowered pressure in the flow between the tarp and the ground. Some vehicle body
components, such as a splitter and a diffuser, can be found under the vehicles body to
help increase the ground effect and improve the downforce of the vehicle. This helps it
travel faster through the corners by increasing the vertical force on the tires.
4

Ground effects in aerospace applications are due to the proximity of the flying
body to the ground. The most important of these effects is the wing in ground (WIG).
This is due to the reduction in lift experienced by an aircraft as it approaches a height of
roughly the aircrafts wingspan above the ground. Those effects increase as the aircraft
approaches the surface, which can lead to loss of control and crashes.

1.3.3. Definition of CFD and CFD History


Computational Fluid Dynamics (sometimes referred to as CFD) is a branch of
fluid mechanics which uses complex algorithms in conjunction with numerical methods
to solve the partial differential equations describing fluid flow. Advances in CFD
software make it possible to perform complex calculations to simulate the interaction of
gases and liquids with each other and geometric surfaces defined by Computer Aided
Design (CAD) software. Yet even with modern high speed computers, only approximate
solutions can be achieved in most cases, particularly for flows involving turbulence and
flow separation around blunt bodies because CFD solution is a numerically based.
CFD originated in the early part of the 20th century, marked by initial attempts to
solve differential equations found in physics and engineering. The main equations
governing fluid flow behavior are the Navier-Stokes equations, developed in the early
part of the 19th century by George Stokes and Claude Navier. Although the NavierStokes equations were a significant development, the analytical mathematical solution of
those equations proved untenable at that time period. This led to the development of a
large number of simplified equations derived from the Navier-Stokes equation for special
cases, which can be tackled analytically using pen and paper or a simple calculator.

However, these special cases were very limited in terms of describing practical
applications.

[36]

The invention of digital computers led to many changes in solving the


complicated Navier Stokes equations. In the late 1940s, John von Neumann led a group
of scientists and engineers to develop modern CFD. The digital computing machines
have the analytical solutions of simplified flow equations with numerical solutions of full
nonlinear flow equations for arbitrary geometries. Modern day CFD uses high-speed
computers to achieve better solutions and improve accuracy of known exact and nonexact solutions to the Navier-Stokes equations such as nonlinear partial differential
equations and turbulence analysis.

[36]

Common CFD codes have a specific structure that revolves around a numerical
method or numerical algorithm able to undertake complex fluid flow studies. Most of the
CFD codes currently on the market have only three basic elements, which divides the
complete simulation to be performed on the specific domain or geometry. The basic three
elements are the following: 1. Pre Processor, where the solution domain is defined and
the mesh is generated; 2.Solver where the flow equations are solved for the previously
defined mesh and domain; and 3. The Post-Processor, where the numerical results are
displayed and analyzed.

1.3.4 Drag Estimation Techniques Using CFD Approach


There are several approaches to calculate the drag on a 3-D geometry using the
CFD approach. Perhaps the most common and widely used approach to finding drag
using CFD is solving the Reynolds Averaged Navier-Stokes (RANS) equations, or the
surface integration of stresses, i.e. near field methods. There are several problems with
6

this approach to solving CFD problems. For the near field method the problem is usually
insufficiently accurate results, for example even if the flow solution is locally accurate in
terms of pressure and velocity profile. As for RANS, the problem is mainly related to the
numerical solution that generates the drag coefficient. A second problem for the RANS is
near field drag computation; it only allows for distinction between pressure and friction
drag. [26]
Due to the mentioned problems above with the RANS methods, the following
approach is used in this thesis to find the drag coefficient of the HPV fairing. This
approach is to use the oblate ellipsoid to determine computational precision of FLUENT
by finding the proper mesh parameters and turbulence model to provide accurate drag
estimates. This approach establishes how fine the mesh should be in order to acquire
proper results for drag forces over the HPV fairing. This mesh incorporates estimation of
the boundary layer thickness to ensure that there are enough points used near the body
surface to accurately predict the velocity gradient within the boundary layer, and the
related friction drag. Using the ellipsoid body geometry, a strategy is developed to
optimize the program settings within the FLUENT solver for effective convergence and
solution accuracy.

1.3.5 Experimental Values of Drag on the Ellipsoid Bodies


An oblate ellipsoid is a disk shaped spheroid where a=b>c, and prolate ellipsoid is
a rugby ball shaped spheroid where a=b<c.
Drag research on oblate ellipsoids and other similar shapes is very limited. There
are only a few real-world examples of such types of bodies. The HPV fairing is being
assumed as a streamlined geometry and the oblate ellipsoid is used to help with the initial
7

setup of the CFD approach. However, there is a lot of literature that discusses drag
information on similar types of bodies, such as prolate ellipsoids and spheroids. This may
be used as a baseline reference for the work being performed in this study. The
information in Figure 1-2 comes from a well-known drag expert, Dr. S.F Hoerner. In his
book Fluid-Dynamic Drag (1965), Hoerner presents the drag coefficient of numerous
shapes such as oblate ellipsoids, prolate ellipsoids, and spheroids in both 2-D and 3-D
flow fields. Figure 1-2 presents the wetted area drag coefficient of an oblate ellipsoid
with different fineness ratios of body of revolution

over a range of Reynolds number

(R e ). The d is the diameter of the ellipsoid at its widest part, and l is the length of the
ellipsoid. The points that are shown in Figure 1-2 are the experimental data that were
found for those bodies, and the dashed lines represent the theoretical drag for fineness
ratio and is given with the following equation.
.

= , +

[6, 11, 12]

+ .

(1-2)

Figure 1-2:
Drag Data on 3-D Bodies of Revolution Aligned Straight-and-Level (Adapted from
Hoerner Fluid Dynamics of Drag, 1965, 6-16)

Figure 1-2 represents the effect of Reynolds number on the drag of the ellipsoid
with different fineness ratios of

. In the laminar region where the Reynolds number is

less than 105 the drag coefficients tend to be higher. When the Reynolds number reaches
between 105 and 106, the boundary layer flow begins to transition from laminar to
turbulent, and a significant drop is seen in the drag coefficient. After the drag reaches its
minimum value, the drag begins to rise slightly as the boundary layer transition point
continues to move forward. Finally, when the Reynolds number reaches 107, the flow is
fully turbulent and the drag starts to decrease again. In reference to Figure 1-2 the higher
the Reynolds numbers, the lower the drag at the fineness ratios. Additionally, the higher
the fineness ratio the lower the drag coefficient will be.

To define the fineness ratio that is used in Figure 1-3 and its relationship to the
friction drag coefficient C f the following equation is employed.

= + .

(1-3)

To find the correct ratio of wetted area to frontal area

for streamline

bodies, the wetted area can be approximated as = (0.7 0.8) ,


where the perimeter is equal to , and the frontal area is equal to

2
4

wetted area to frontal area is equal to:

This expression is then substituted into equation 1-3 to find the


area coefficient and curve fit for Figure 1-3 as derived by Hoerner.

0.5

= 3 + 4.5

+ 21

. The ratio of

for the frontal

(1-4)

Figure 1-3:
Drag coefficient of streamlined bodies as a function of their thickness ratio (Adapted
from Hoerner Fluid Dynamics of Drag, 1965, 6-19)
10

Figure 1-3 illustrates the drag coefficients based on frontal area of streamline
bodies as a function of their fineness ratio, the points in Figure 1-3 are the experimental
data and the dashed lines are evaluated from equation 1-4. When the fineness ratio
increases, the drag coefficient also increases expect for low fineness ratios.
The drag coefficient for the HPV fairing based on its frontal fineness ratio of 3.53
is between 0.02 and 0.065 for Reynolds numbers 105-107. This was found using Figure 13 and equation 1-4.
It is difficult to isolate the critical Reynolds number on the oblate ellipsoid where
the transition will occur from laminar to turbulent flow with estimated Reynolds numbers
[8]

from 500 to 600 thousand for that geometry. Figure 1-4

shows the wetted area drag

coefficient for the = . prolate spheroid for several different surface roughnesses. The

roughness has an enormous effect on the drag coefficient in the low Reynolds numbers.

This is because the flow is not fully developed and this adds to the total skin friction
coefficient as illustrated in figure 1-4. During Dr. Dresss study the critical Reynolds
number reached about 800 thousand where the transition from laminar to turbulent region
occurs, and the minimum drag coefficient happened at a Reynolds number of almost 1.2
million for a fine grit of 80. The different types of runs show the effect of skin roughness
from laminar to turbulent flow, and the effect on the wetted drag.

11

Figure 1-4:
Drag Data from a = . Prolate Spheroid Aligned Straight-and-Level free transition

is the base run, 80 is the fine grit, and 40 is the rough grit (Adapted from Dress, NASA
Technical Paper 2895 1989, 29)

1.3.6.

Drag Values on Variation With Ground Clearance


Once the potential of using aerodynamic downforce in automotive racing

applications was realized, many teams started to experiment with other methods to
increase aerodynamic downforce other than simply attaching inverted wings. It was
found that with a larger underbody area of the vehicle, significant levels of downforce
could be generated. This kind of effect was first seen in 1935 in the racing circuit with
early wing prototypes used in ground effect models.

12

[13]

Figure 1-5 illustrates a basic principle of ground effect on typical car shapes
represented by an oblate ellipsoid and half streamlined body. However, to understand
ground effect the nature of the flow under the vehicle must be considered. The top part of
the Figure, shows an oblate ellipsoid that is approaching the ground. The flow under the
oblate ellipsoid and the downforce (C L ) are increasing as distance to the ground reduces
and creates low pressure. If one looks at the bottom part of the Figure and closely
examines the half streamlined body, the drag coefficient is seen to be nearly the same as
the oblate ellipsoid. The lift force is opposite due to the reduced flow under the body,
with the result of increased lift due the reduced ground clearance. In both Figures the

transition to significant ground effect starts to occur at < 0.05. However, this only

applies to these specific geometries. The transition point can shift to either left or right
depending on the fineness ration and overall shape of the geometry.
There are several options for the car body shape to generate lower pressure under
the body. Option one is to streamline the underbody to create low pressure. Option two is
to create a seal between the underbody of the car and the ground and only leave the rear
portion of the car open. Then the low pressure behind the car would dictate the pressure
under the car.

[14, 15]

13

Figure 1-5:
Effect of ground Proximity on the lift and drag of two streamline bodies (Adapted from Race
Car Aerodynamics by Joseph Katz 1995)

14

1.4.

HPV Fairing Geometry Description


Figures 1-2 and 1-3 are used as a reference to estimate the expected drag

coefficient for the HPV fairing. If one assumes the HPV fairing is a body of revolution
then the oblate ellipsoid can be used as a computational precision tool for the mesh setup,
turbulent model selection, and optimize FLUENT parameters. To apply Figures 1-2 and
1-3 one needs to estimate an equivalent fineness ratio for the HPV fairing, and a range of
drag values can then be estimated for the HPV fairing in freestream flow. This is used as
a benchmark for the HPV fairing analysis.
Figure 1-6 shows the dimensions of the HPV fairing; this data can then be used to
find the fineness ratio based on the height of the HPV fairing which is equal to 3.53 for
half of a body of revolution. However, because the HPV fairing is assumed to be a body

of revolution the height needs to be doubled to get the correct fineness ratio = 1.76. the

resulting wetted area drag coefficient value for


C D,Wetted =0.009 and C D,surface area =0.091.

= 3 106 is approximately

Figure 1-6:
Dimensions of the HPV fairing from SolidWorks 2010 where l= 98.93 inches, h=d= 28.03
inches

15

1.5.

Organization of the Thesis


The remainder of this thesis will be organized as follows. Chapter 2 describes the

model design and importation of the model into ANSYS WORKBENCH and fluid
volume extraction. It also explains how to import the model in to ANSYS ICEM and the
mesh setup and creation. Lastly, it will be explained how to import the mesh from
ANSYS ICEM to ANSYS FLUENT.
Chapter 3 explains how to operate FLUENT using ANSYS WORKBENCH and
apply FLUENT setups as an initial setup, materials for the fluid and geometry, dynamic
mesh, and boundary conditions. It will demonstrate how to use FLUENT to generate
numerical and graphical solutions for the HPV fairing geometry with different ground
clearances ranging from 3 inches to 18 inches away from the ground.
Chapter 4 presents the results of the baseline solution of the oblate ellipsoid with

= 2&4 and results for the flat plate. This chapter also compares the CFD results of the

baseline solution to the results found in Chapter 6 in Fluid Dynamics of Drag by


Hoerner.

[11]

Chapter 5 presents the results of the HPV fairing with different ground clearances
ranging from 3 inches to 18 inches away from the ground. Then the results from the HPV
fairing CFD analysis are compared to the benchmark results (freestream and 30 inch
ground clearance). In addition, the results for drag and lift are discussed, and calculations
of discretization error are presented. Then the final part of Chapter 5 will include the
trade-off study regarding the optimum vehicle height while considering both vehicle
stability and aerodynamic drag.
16

Chapter 6 is the conclusion and the summarization of the study. It is based on the
results shown in Chapters 4 and 5. References and an appendix follow the conclusion.

17

Chapter 2: Importation of Solid Model into ANSYS and Mesh Definition


2.1.Meshing and Preprocessing
The pre-processing of a CFD procedure consists of several inputs for the flow
problem that are done by the user in CFD software. For this study the pre-processing
software is ANSYS ICEM CFD, and the solver software is ANSYS FLUENT. The inputs
are then transferred into a form made suitable for use by the solver. The pre-processor is
the main connection between the CFD solver and the user. The user has to complete
several significant steps in the pre-processing stage of the CFD problem. A schematic of
the process is shown in Figure 2-1.The following definition, gives a brief explanation of
these steps.
1. Define the geometry of interest: This step uses ANSYS DesignModeler CAD
software within ANSYS WORKBENCH to help design and model the topology of
the fluid flow domain inside or outside the geometry. This domain is defined and
optimized for the best CFD results.
2. When the geometry preparation is defined within the pre-processor software, the fluid
domain and every surface affected by the fluid is then also defined. Each fluid and
surface has its own distinct property; those properties are used in the CFD process
and must be defined at this stage. The output of the DesignModeler software is a
xxxx.agdb file.
3. Meshing is the third step. Because the CFD process uses a finite volume method, the
domain of interest has to be divided into structured and unstructured elements. All the
elements are connected to each other through nodes to and from the flow domain. For
this study ANSYS ICEM CFD software is used to create the mesh in the form of a
18

xxxx.mesh file. The quality of the mesh contributes to the accuracy of the final
results.
4. Definition of boundary conditions is the final step at the pre-processing stage. Each
CFD domain needs an initial condition to begin calculations, which is defined by the
users input. In addition, the CFD code implements the boundary conditions at a
specific locations.
The following few sections will explain these four steps in complete detail and
explain how to use ANSYS 12.1 for external flow problems. Lastly, Figure 2-1 illustrates
how the files from the different software packages move through the overall solution
process.

.agdb

SolidWorks

ANSYS
ICEM
CFD

.mesh

.SLDPRT
ANSYS
DesignModeler

ANSYS
FLUENT

ANSYS
WORKBENCH
Figure 2-1: Block Diagram illustrates where each file type goes to

19

.wbpj

2.2. Modeling of the HPV Fairing and the Ellipsoid Geometries in SolidWorks
All of the solid models that were used in this study were designed and drafted
using SolidWorks Computer-Aided Design (CAD) software, using inches for dimensions.
The fairing was originally designed and modeled by the 2009-2010 California State
University Northridge (CSUN) Human Powered Vehicle (HPV) Team for their
competition in April 2010. An ellipsoid model was also designed to represent a simpler
geometry and was used as the baseline for this thesis. The ellipsoid model establishes the
mesh fineness requirements to acquire good results for the drag force, based on
comparison with published results from Fluid Dynamics Drag by Hoerner data.

[11]

The modeling of the ellipsoid geometry in SolidWorks was a little challenging,


because the ellipsoid had to represent the fairing shape as closely as possible. The
ellipsoid was created using the lofted boss/base tool in SolidWorks. However, before that
could be done, several planes were created so that a 2-D ellipse could be drawn on each
plane with different chord lengths A and B. This is illustrated in Figure 2-2.

Figure 2-2:
Representation of an ellipse geometry B=99in and A=49.5in

After all of the 2-D schematic geometries were drawn, the lofted boss/base tool
was used to create the 3-D ellipsoid body that can be seen in Figure 2-3. The ellipsoid
20

model dimensions are: the chord length (l) is 99 inches; height(x) is 49.5 inches, and the
diameter (d) of the ellipsoid is 24.75 inches. The fineness ratio of
as

99

49.5

is then can be found

= 2. This ratio is then used to find the drag of a non-oblate ellipsoid body.

Additionally, another ellipsoid was created in SolidWorks with a fineness ratio of

= 4,

and was used as a baseline test in FLUENT. Additional comparisons were made with a

flat plate geometry which is useful because the drag force on a flat plate is completely
due to surface stresses.
There are a few reasons why two oblate ellipsoids are used to calibrate FLUENT
and set correct mesh parameters for the HPV fairing. The first reason is to match the
results from FLUENT runs to the known results from Fluid Dynamics Drag by Hoerner.
The second reason is to find the limitation of FLUENT on predicting drag on similar
geometries with different fineness ratios, as the flow behaves differently for a Falter
shape. Generally, a smaller

ratio will have a larger contribution of pressure forces to

the overall drag, especially if the boundary layer separates on the rear portion of the
body.

Figure 2-3:
3-d Ellipsoid body from SolidWorks

21

After the models were created and saved in SolidWorks, one needed to import
those models into ANSYS 12.1 for geometry calibration and model clean up before the
models were meshed and used within ANSYS FLUENT.

2.3. Importing Model into ANSYS WORKBENCH from SolidWorks


ANSYS WORKBENCH is a Computer Aided Engineering (CAE) software
package that is used in engineering simulation and analysis. It is an innovative project
organizer that ties together the entire simulation process. It helps the user go through
several complex studies at once with drag and drop menus. It also has powerful user
controls, automated meshing abilities, project level update mechanisms, and integrated
optimization tools, which enable complex simulation and product optimization.

[40, 37]

The next few Figures show a step by step explanation process to import any
SolidWorks model into ANSYS WORKBENCH, and clean up the geometry so it can be
properly meshed. Figure 2-4 shows how to load the geometry in ANSYS
WORKBENCH. In order to load the SolidWorks model in ANSYS WORKBENCH, the
user first has to open ANSYS WORKBENCH, then go to the component systems and
select Geometry (A). Then the geometry tab is placed on the main WORKBENCH
screen, and it then becomes a cell. In order to load the geometry, the user must right-click
the Geometry..? tab, and then scroll down until import geometry has been reached.
After left-clicking on this item, a new window will open. Then user must left-click
browse tab and load the specific geometry (B) to be modified.
After the geometry is loaded into the WORKBENCH, the user must double click
with the left mouse button on the geometry cell number 2, and ANSYS DesignModeler

22

will load. The user then is able to clean, modify, edit and fix the geometry so a better
mesh can be created for future analysis of the model. This is explained in Section 2.3.1.

A
B

Figure 2-4:
ANSYS WORKBENCH front screen; A- geometry is selected first; B-geometry cell
where geometry is going to be imported

23

2.3.1. Extracting a Fluid Volume for the Models


The next few Figures will show step by step how to extract the fluid volume
around the imported geometry. The fluid volume must be extracted because one must
correctly define the volume that is being occupied by the fluid around a specific solid
model.

Figure 2-5:
Ellipsoid model with in ANSYS DesignModeler and the selection of the external flow.

Figure 2-5 illustrates how once the geometry is loaded into ANSYS
DesignModeler the user can then begin to select what kind of fluid volume to apply to the
specific model, such as internal or external fluid volume. For this study an external fluid
volume is being used. This is because the imported geometry represents a solid body and
the air flow is external to the body surface.

24

Figure 2-6:
Selection of shape and cushion type

Figure 2-6 illustrates the shape and the cushion size of the fluid volume enclosure.
The cushion size is also known as the domain size. For this study the shape of the fluid
volume is the box shape, since it is convenient for generating the mesh around the solid
body. Since the CFD process is a numerical approximation approach that uses the finite
volume method to solve the NavierStokes equations, the fluid volume domain is going
to be composed of an Octree Mesh, sometimes referred to as an unstructured mesh. In
order to create the fluid volume domain, the user must set the cushion size and select
either uniform or non uniform size. For this thesis the non-uniform cushion size will be
used on all the models. This is done to make a more efficient study that does not require a
large quantity of computing power.
The ellipsoid model was run in freestream condition without any ground plane
representation. The HPV fairing simulation consisted of eight different cases. The first
two cases are set as benchmarks, where one is in freestream condition and the other one
simulation a ground clearance of 30 inches. The other six cases will simulate the HPV
fairing with ground clearances ranging from 3 to 18.
25

Figure 2-7:
Generated enclosure for the oblate ellipsoid in freestream

Figure 2-8:
Editing of the enclosure based on symmetry

Figures 2-7 and 2-8 illustrate the generated fluid volume enclosure for the solid
model, and the editing process for the fluid enclose based on model symmetry about the
XY plane. This makes the computation more efficient because it only has to analyze half

26

of the model to achieve the same results. In order to create the symmetric model, the user
must right click on the Enclosure tab in the tree outline, and then select the edit
selection tab. After the user has selected the previous command, the model enclosure can
then be edited to the users specifications and the correct symmetry plane.
The user can then select up to three planes of symmetry. As mentioned earlier this
model is only symmetric to one plane, the XY plane. In order to select the symmetry
plane, the user must left click on the not selected tab and then the user must select the
corresponding plane from the tree outline, then press apply. In order to generate the new
model, the user must press the Generate tab to create the symmetric model about the
XY plane. This is illustrated in Figure 2-9 where one can see the selection of the total
number of planes that can be used at the same time, and the symmetry plane selection.

Figure 2-9:
Selection of symmetry planes. For this study it is the XY plane.

27

Y =3x

X =6x

X =3X

Z=3x

Y =3x

Figure 2-10:
Fluid volume for the ellipsoid model

Figures 2-10 and 2-11 show the final view of the oblate ellipsoids model and
fluid enclosure, and the HPV fairing within the non-uniform fluid volume box. The
oblate ellipsoid fluid volume box is X+=Y+=Y-=Z=3 times chord length, and X-=6 times
chord length. The fairing fluid volume box is X+=Y+=Z=3 times chord length, X-=6 times
chord length, Y-=3 to 18 inches for the test cases, and for the benchmarks it is 30 inches
and 297 inches. The domain size was selected to help decrease the total computing power
while maintaining accuracy. The optimal domain size for a wing was found by Amir
Mohammadi in his thesis and this data is being used as a reference for the domain size
used here.

[21]

Before the mesh can be created, the model needs to be exported as an


xxxxxx.agdb file. In order to save the ANSYS DesignModeler file, the user must do the
following steps; File>Export> xxxxxx.agdb> then Save. Once the file is saved, it then
can be opened by ANSYS ICEM CFD, and a proper mesh can be applied to the solid
model and the fluid volume box.
28

In addition to creating the fluid volume, naming the surfaces that represent the
boundary conditions will help later with ANSYS FLUENT setup and the meshing
process in ANSYS ICEM CFD. In order to name the different surfaces, the user must
right click on the surface and then click edit to name the surface. For the oblate ellipsoid
and the HPV fairing model, the surfaces that are created are the inlet velocity, outlet,
boundary volume box, and symmetry plane. The boundary volume box for the oblate
ellipsoid is made out of three surfaces that surround the geometry. However, for the
fairing the bottom surface is named ground plane and the volume box is made only of
two adjacent surfaces. This is illustrated in Figure 2-12 and Table 2-1.

Y =3X

X =6X

X =3X
-

Z=3X

Y =3 to 18 inches

Figure 2-11:
Fluid volume for the Fairing model

29

Plane Name

Surface Name For Ellipsoid

Surface Name For


HPV

Right (YZ plane @ X+)

Velocity Inlet

Velocity Inlet

Left (YZ plane @ X-)

Outflow

Outflow

Fluid Volume Box

Fluid Volume Box

Fluid Volume Box

Ground plane

Fluid Volume Box

Fluid Volume Box

Symmetry Plane

Symmetry Plane

Top (XZ plane @ Y+)


Bottom (XZ plane @Y-)
Far side (XY plane @ Z+)
Symmetry (XY plane @Z-)

Table 2-1:
Surface names for ellipsoid and HPV Fairing

Fluid
Volume
Box

Outflow

Velocity
Inlet

Ground
plane

Symmetry

Figure 2-12:
Plane location and names

30

2.3.2. Opening the Models in ANSYS ICEM CFD


Before the meshing procedure can begin, the file that was saved by
DesignModeler must be opened in ANSYS ICEM CFD. In order to do that, the user must
do the following steps; File>WORKBENCH Reader>select xxxxxx.agdb file> then
Open. Prior to the file being completely loaded into ANSYS ICEM CFD, the user has to
go to the scroll down menu below and select the options that are illustrated in Figure 213. Then the user must press apply.

Figure 2-13:
Importing an xxxxxx.agdb file into ANSYS ICEM CFD CFD (A). Opening the
xxxxxx.agdb in ANSYS ICEM CFD CFD (B)

31

2.3.3.Preparing the Geometry for Meshing

Figure 2-14:
Extracting the feature curve from the symmetry plane

Select those
locations for the
fluid volume area

Figure 2-15:
Demonstration the correct location

32

Figure 2-14 illustrates how to prepare the geometry that was loaded into ANSYS
ICEM CFD so that the correct mesh and grid can be generated. In order to extract the
curves from the surface, the user must do the following steps: Geometry tab > Create/
Modify Curve icon> Extract Curves from Surfaces icon, then select the surface on the
screen. The user has to click on the glass icon to select all appropriate visible objects, or
use the following shortcut key v. The plane that is selected for this study is the
symmetry plane. After all the correct surfaces are selected, the user must click apply or
OK.
Following the Extract Curves procedure, the body for the fluid has to be created.
In order to do that, the user must start with the Geometry tab again, and then the user
must click the Create Body icon. Following that, name the part as the fluid name; any
name can be used to name the region. For this study the name that is used is Fluid
Volume. In order to name the fluid region, the user must select Centroid of 2 points for
the location and the Material Point icon to select the location of the fluid volume. Then
the user must click the two screen locations to select the fluid body region as
demonstrated in Figure 2-15. Following that, the user must click OK to finish creating
the fluid volume area and proceed to the meshing setup. In addition, the user must create
parts from the Subsets by selecting the inlet velocity, outlet, and the fluid volume
boundary, and then right click on the Subsets to create parts. These names, are used
when meshing in ANSYS ICEM CFD, and setting the boundary conditions and
parameters in ANSYS FLUENT.

33

2.3.4. Generating the Initial Mesh Using the Octree Mesh Approach and Applying
the Correct Mesh Size
The strategy that is used for the mesh process is to have a prismatic or structured
mesh around the solid model and then transition to an unstructured mesh. The prismatic
mesh represents the boundary layer and is defined as a stair step mesh to decrease the
required computing power. The height and the mesh density of the prismatic layer was
set to represent the estimated boundary layer thickness around the solid models, i.e.
oblate ellipsoids, flat plate and HPV fairing. Then the prismatic mesh transitions to an
unstructured mesh to create a hybrid mesh around the solid model and inside the fluid
region.
Assigning the correct mesh for each model was a trial and error method. The
reason behind this is that each model used slightly different mesh parameters, and it also
varied from robust to fine mesh. It also depended on the size and shape of the geometry.
The Scale Factor multiplies other mesh parameters to globally scale the model, for
example if a Max Element Size of a given entity is 64 units and the Scale Factor is 0.3
units, then the actual maximum element size will be 64 0.3 = 19.2 . After

countless tries, the correct scale factor was found to be approximately 0.3 for all the
models. For that reason, all the models used a proper mesh for balancing accuracy with

computed memory requirements.


The maximum element size that was selected ranged from 64-128. This value was
selected due to the fact that an Octree Mesh scales by a power of two, and the Octree
algorithm is limited to datasets of resolution of power of two. For that reason our values
range from 26-27 (or 64-128). This is very important because all other values that will be
input into the maximum scale factor will be rounded off to the closest power of two. In
34

order to set the parameters, the user must select Mesh tab> Global Mesh Setup icon >
Global Mesh Size. After the correct input is input the user must click apply/ok. This is
demonstrated in Figure 2-16. Lastly, the general grid topology will be talked in chapter 3.

Figure 2-16:
Meshing sizing with ellipsoid of ratio l/d=2

After the meshing sizing is completed, the user must select the Part Mesh Setup
icon. This icon is selected in the Mesh tab area to specify the mesh parameters. In order
to create the prism mesh, the user must first select the prism option in the mesh parameter
area, only for the solid model and the symmetry plane. The prism height is set to 0.1-0.2,
depending on the model, so it can build the correct boundary layer as learned in ME692.
For the ellipsoid and fairing geometry surfaces the maximum size is set in the range of
2.5-3; this creates a proper surface mesh for the solid geometry. Also the user needs to
input at least 90 for number of prism layers of to be created, and a height ratio of 1.06-1.1
for the growth factor. This corresponds to the maximum thickness () in the turbulent
boundary layer, which is approximately 2 inches. This number was found using the
35

calculations that can be seen in appendix A for the boundary layer thickness for the
laminar, turbulent, and transition layers on a flat plate with a length equal to that of the
ellipsoid and the fairing models.
For the fluid volume box (inlet velocity, outlet, symmetry and open domain) the
maximum size is set to 64 to allow create an appropriate volume mesh. After the mesh
parameter setting are complete, the user must press apply. This is shown in Figure 2-17
for the ellipsoid and HPV fairing models.

Figure 2-17:
Mesh parameters step for the ellipsoid

Figure 2-18:
Mesh density box setup (a). Shifting of mesh density box to refine wake region (b)

The density box is created to represent the wake region of recirculation flow
immediately behind the model. The wake region is chaotic due to boundary layer
separation on the rear portion of the body. The density box allows local control over the
mesh density in the wake region to correctly represent the flow.
36

In order to create the density box that represents the wake region, the user must
first select the Mesh tab> Create Mesh Density icon, then select the size of the density
box. For this study the size was selected at 32, and the ratio and width were left at zero.
The user then must select the density location as an entity. After the density box is
selected, the user must click OK to generate it. Note that at this point the box surrounds
the solid body. In order to shift the density box to the expected wake region location, the
user must click Geometry tab>Transform Geometry icon >Translate Geometry icon,
then select the density box and keep the translation method as explicit. Before the density
can be shifted the model needs to be measured by the Measure Distance feature.
Following that the density box is shifted by half of the model length. In this study the
model was 99 inches long so the density box was shifted 44.5 inches in the negative X
direction to represent the True Wake region. This is illustrated in Figure 2-18.
Following the completion of the creation and shifting of the density box, to
generate the mesh, the user must first click the Mesh Tab>Compute Mesh icon, then
the user must select the Create a Prism Layers and click Compute, as Figure 2-19
illustrates. Following that another mesh has to be defined to refine the present mesh of
the model that can be correctly analyzed within ANSYS FLUENT. This is the Delaunay
mesh step, and it will be discussed later in the chapter. The reason why an Octree Mesh
was used as opposed to a Delaunay Mesh is to minimize the numerical error as much as
possible. This also helps to minimize the total computing power needed to create a solid
mesh.

[39]

37

Figure 2-19:
Computing the initial mesh

Figure 2-20 illustrates the cut plane that allows the examination of the prism
layers in the mesh around the solid model. Please note that the prism height floats, as the
height was initially set to 0.05-0.1. These numbers illustrate that the first few prism layers
start growing very slowly and there after grow exponentially. The variation in layer
thickness (float) is not significant for the model because the surface mesh size is
relatively uniform. The mesh density near the solid body does not vary with axial
position as defined in ANSYS ICEM CFD, note that the mesh is adjusted during the
analysis in ANSYS FLUENT with mesh adaption.

38

Figure 2-20:
Mesh analysis using a cut plane in the XY plane (A); YZ plane (B); XZ plane (C)

The mesh process is completed by performing a check done on the mesh to find
any errors that may cause problems during the analysis in FLUENT. In order to check the
mesh, the user must do the following steps; Edit Mesh tab > Check Mesh tab. The
user needs to keep the default settings and then click OK.

39

2.3.5. Generating the Tetra/Prism Mesh Using the Delaunay Mesh Approach
Once an Octree Mesh has been checked and no errors have been found, the
Delaunay Mesh can be generated. The Delaunay Mesh more efficiently fills the volume,
and it has a smoother volume transition. This kind of mesh works a lot better with
FLUENT to help calculate better results for drag for all the models according to the
ANSYS ICEM CFD user manual.

[39]

Figure 2-21 displays the steps to generate the

Delaunay Mesh within ANSYS ICEM CFD.


In order to generate the Delaunay Mesh, the user must do the following steps:
click on Mesh tab>Global Mesh Setup>Volume Meshing Parameters, and select the
Delaunay option from the drop down menu. The user must enter a scale factor of 1.2,
memory scaling factor of 1 and the Delaunay Scheme must be T-Grid according to the
ANSYS ICEM CFD user manual.

[39]

After all the correct options have been selected, the

user must click Apply. In order to start the computing process, the user must click on
the Compute Mesh icon and select the Delaunay method from the drop-down menu, and
then disable the Create Prism Layers option. The user must make sure that the Existing
Mesh option is selected from the drop-down menu because the mesh is generated based
on the Octree Mesh; then finally click Compute
.

40

Figure 2-21:
Delaunay Mesh Setup

2.3.6. Smoothing the Mesh to Improve Quality


The smoothing of the mesh is done to improve its quality. The smoothing
approach involves the initial smoothing of the interior elements without adjusting the
prism elements. After the initial smoothing is complete, the prism elements then will be
smoothed by themselves.
In order to smooth the mesh, the user must click the Edit Mesh tab > Smooth
Mesh Globally tab. To smooth the mesh that was generated using the Delaunay
Approach the user first has to smooth the interior elements without touching the prism
elements. This is done by opening the Smooth Elements Globally control panel. The
first step that the user must do in this process is to set the number of smoothing iterations;
this number was set to 25. The second step is to enter the Up to Value; this value
41

specifies the quality level up to which the program will attempt to smooth the mesh. It
was set to 0.5 this was based on ANSYS ICEM CFD settings.

[39]

Then for the criterion

the user must select the quality option from the drop down menu. Lastly the user must set
all the elements to get smooth except for PENTA_6 which was set to freeze.
The reason why PENTA_6 was set to freeze is because it is a five sided element
with six nodes as a prism element. These elements are usually perfect, but they may be
damaged by the smoother as it adjusts to optimize the nearby tetra elements. By selecting
the freeze option in the Smooth Mesh type for the PENTA_6 elements, it protects them
from being damaged. When smoothing those kinds of elements the values for the Up to
Value should be reduced to 0.01 so only the worst of the PENTA_6 elements are
adjusted, and the number of smoothing iterations should be dropped to 2. Figure 2-22
illustrates how the smoothing step is setup and the quality Histogram for the mesh
elements.

[39]

Figure 2-22
Mesh smoothing setup and quality histogram.

42

2.4.

Exporting the Mesh into ANSYS FLUENT


When the mesh process is finally completed, checked, and smoothed, the user has

to then save the project and transfer the mesh into ANSYS FLUENT. This procedure
applies to all the models for this thesis and can be used as a general guideline for future
CFD projects.
There are several steps in this procedure of transferring the mesh file from
ANSYS ICEM CFD to FLUENT. The first step is to save the ICEM project by clicking
on File>Save Project As, which creates a xxxx.uns file. The second step is to go to the
Output tab and select the red tool box (Select Solver). After the Select Solver is clicked
a menu will appear on the screen with two drop down lists. The first list is Output Solver;
the user must select the FLUENT_V6 option in order to produce a mesh file that is
compatible with FLUENT. The second drop down list is the Common Structural Solver;
the user must select ANSYS option, and then click Apply; as illustrated in Figure 2-23.

Figure 2-23:
Output step and solver selection

43

Following the Output Solver and the Common Structual Solver selection, the user
then can apply the boundary conditions to mesh. The boundary conditions are located in
the Output tab, where the user can apply the boundary conditions and check that all the
surfaces are defined and represented correctly. This is illustrated in Figure 2-24.

Figure 2-24:
Boundary condition step.

After all the the above steps are completed the user can then write the input file
for ANSYS FLUENT. This is done in the Output tab once again. In order to write the
mesh as a FLUENT compatable file, the user must select the Write Input tab. First the
correct ANSYS.uns file for the project, (that was saved in the first step) must be opened.
Then the FLUENT_V6 window will appear. Following the windows appearance a name
for the file must be entered in the Output File line. All other options can remain as the
defult values; the step is completed by clicking Done. This is all illustreted in Figures 225 and 2-26 below.

44

Figure 2-25:
Opening of the ANSYS .uns File

Figure 2-26:
Fluent_V6 window that appears after the ANSYS .uns file is selected.

After the mesh is saved as a FLUENT file (.msh file) the user then can close
ANSYS ICEM CFD, and open ANSYS WORKBENCH. In order to load the mesh into
45

FLUENT, the user must select the mesh option from the component systems list and drag
it to the WORKBENCH. The same thing is then done for the FLUENT option. After the
two boxes appear on the WORKBENCH, the user must right click on the mesh cell in the
mesh box and load the FLUENT mesh, as illustrated in Figure 2-27.
After the mesh has been loaded in the ANSYS WORKBENCH, it then can be
loaded in FLUENT. This is done by dragging the Mesh cell from the Mesh box to Setup
cell in FLUENT Box.

Figure 2-27:
Loading of the .msh file in ANSYS WORKBENCH

46

Chapter 3: FLUENT Setup and Application of Spalart-Allmaras Turbulence Model

3.1.

Background Information in Computational Software and Methodology


The major reason behind the growth of CFD usage in various industries is due to

its accuracy, reliability, and replacement for running experimental tests. There is also
much more advanced computing technology available today for much less cost than
running a physical experiment, which may require major equipment such as a wind
tunnel. This kind of software is capable of solving large two and three dimensional
problems numerically in a short period of time.
The accuracy and reliability of a CFD simulation depends on the numerical
algorithms employed by the software. This means selecting the appropriate options such
as a turbulence model, appropriate spatial and temporal discretization scheme, and
correct computational grid topology. The grid topology can have significant weight on
the final results of the CFD simulation. Each one of the options mentioned earlier can
have either a positive or negative effect on the simulation.
With reference to the grid topology, structured grids are more common, preferred,
and efficient in the boundary layer region along the model surface for the simulation of
the flat-plate, oblate ellipsoid and HPV fairing. In addition, structured grids allow more
efficient computations and parallelization. However, an unstructured grid requires less
grid points outside the boundary layer region. Considering the oblate ellipsoid geometry,
the unstructured grid was a lot easier to generate; it also adapted to the flow gradients
more easily. However, the structured grid was much harder to generate around the model
within the boundary layer. The reason why all the models use hybrid grids is to simplify
the mesh creation and provide accurate and reliable results.
47

FLUENT is a finite-volume solver that is based on the full Navier-Strokes


equations with a Blasius assumption for turbulence. FLUENT works on structured and
unstructured grids. As noted above, the mesh for each model is composed of both kinds
of grids. Various grids were examined in order to find the optimum size grid to use for
this study. In the thesis Computation of Flow Over a High Performance by Amir
Mohammadi,

[21]

grid optimization was considered, and some of those findings, have

been used here.


The following section discuss the way FLUENT solves the grid and provides the
user with the proper results. FLUENT uses cell faces to integrate for a solution, since the
software must handle both structured and hybrid meshes. The hybrid mesh contains many
different types of cells such as TETRA_4 (Tetrahedral), TRI_3 (Triangles), PENTA_6
(Prisms), QUAD_4 (Quadrilateral) and PYRA_5 (Pyramids) cells. The structured mesh is
a uniform mesh, composed entirely of QUAD-4 cell.

3.2. Turbulence Model


Turbulence modeling is the construction and use of a model such as SpalartAllmaras (SA), k-epsilon (k-), or k-omega (k-) to predict the effects of turbulence
around or inside blunt objects.

[33]

Averaging is used to simplify the solution of the

governing equations of turbulence; hence the models are required to represent different
scales of the flow that are not resolved.
Consideration of turbulent flows phenomena includes transport properties,
boundary layer separation, and other major phenomena; because of this, the most recent
work focuses on different types of turbulent models that consist of one or two equation

48

models. For instance, examples of two-equation models are the k- and k- models, and
the most popular one-equation model is the SA model. For this thesis, the SA model is
being used because of its strong performance in the baseline studies versus the k- and k turbulent models.
3.2.1. Spalart-Allmaras Turbulence Model
The SA model was developed in 1992 by Dr. Steven R. Allmaras and Dr. P.R
Spalart. The SA model is an approach for modeling different types of turbulent flows,
specifically aerodynamics flows with a high Reynolds number. This model is basically a
transport equation for the eddy viscosity , or a parameter that is proportional to the
turbulent viscosity. The main idea that Spalart and Allmaras used to develop this model
was very similar to the Nee & Kovasznay (NK) model, which was developed in 1969,
and more recently the Baldwin & Barth (BB) in 1990. However, all one-equation models
.[42]

have been based on the turbulent kinetic energy equation

It was discovered during the preliminary and baseline tests on the flat and
ellipsoid models that the SA model provided better results for drag forces and prediction
of flow separation, compared to other options such as the k- and k- models. Due to its
performance during these tests for different flow conditions, the SA model was selected
as the main turbulent model for this thesis. As noted above the SA model employs only
one-equation, which is a partial differential equation for the modified eddy viscosity. The
basic equation is setup as:

+ ( ) = +
49

(3-1)

This can be written as:

= 1 (1 2 ) 1

(3-2)

2 +

( + )

Equation 3-2 can be simplified and the term by term explanation will be given
over the next few paragraphs.

[
]

(3-3)

+ ( + )
+ 2

Or
=

Rate of change of
viscosity
parameter

+ [ () + cb2 ()2 ] [
]

Transport
of by
convection

Or in words:

Rate of
production
of

(3-4)

Transport of
by turbulent
diffusion

Rate of
dissipation
of

The production, diffusion, and destruction terms that were defined in the SA
model were based on the NK model. The production term defined by NK was based on a
statement that was made by Nee & Kovasznay about what defines eddy viscosity and
turbulent flow. The eddy viscosity can be regarded as the ability of turbulent flow to
transport momentum. The ability must be directly related to the general level of
activity, and therefore, to the turbulent energy".

50

[28]

Based on the above argument, NK defines the production term analogous to the
production of turbulent energy. Based on this assumption, NK then assumed that the

production term must increase monotonically with magnitude of the mean vorticity

and the increase of the total viscosity.

The SA model is slightly different in defining the production term in terms of its
consideration of the appropriate form of mean vorticity. Since the NK model focuses on

the simulation of the turbulent shear flow, then the mean vorticity form of was the

best choice. However the SA models emphasis is on high Reynolds number

aerodynamic flow in which turbulence is found only where the vorticity is located.
Consequently the SA model uses only magnitude of the vorticity.
The diffusion term that was defined in NK used a general definition of diffusion
of a scalar F based on the general diffusion equation:
=

(3-5)

Here is the flux of F due to diffusion and it can be rewritten as = ,

where D F is the coefficient of diffusion. In addition to the diffusion assumption by NK,


they also considered the total viscosity = + as a portable quantity, where is the

molecular viscosity and is the eddy viscosity. NK also assumed that turbulent motion
diffuses by itself; for that reason the coefficient of diffusion is assumed to be D n =n, and
henceforth the turbulent Prandtl and Schmidt numbers are approximately one and

1.

Based on all the above NK assumptions for the diffusion term, the equation is given as:
= (nn)
51

(3-6)

The SA model is still slightly different than the NK model for the diffusion term.
SA considers the general diffusion operator as ([ ]), where is the eddy
viscosity and is the Prandtl number. In the SA model, the molecular viscosity does not

play a major role, and the Prandtl number is still about one. The main difference between
the SA and NK models comes in the conservation of the integral. Spalart and Allmaras

pointed out that manipulation of two-equation models such as the k- model often brings
out diffusion terms that are not conserved. For example, if a cross product of k is
calculated, a non conservative diffusion term will then be allowed in the equation.

[28, 42]

Lastly, the destruction term in the SA model is very similar to the NK approach.
NK again uses the same assumptions as the production term for the eddy viscosity to
construct the destruction term. NK states that the rate of decay of the energy of highintensity uniform turbulence is a very rough approximation, and it is inversely
proportional to the square of the energy:

= (
2 )2

(3-7)

Separating the terms and then integrating both sides will then get the decay law:

2 1

(3-8)

Since Equation (3-1) considers the quantity F to be the total viscosity n, if the
production and the diffusion terms are removed from Equation (3-1), it will then reduce
to:

=
52

(3-9)

Based on the NK assumption of similar behavior of total turbulent energy and


viscosity, it can be assumed that 2 . Finally, based on dimensional
analysis, the final form of the destruction term is given as:

( )

(3-10)

The term B is a universal constant for the turbulence production and L is the
characteristic length. The L term was introduced in order to make B a non-dimensional
term. Usually L is a function of y, but in this area of the outer edge of the turbulent flow,
L is assumed to be equal to the boundary layer thickness (). However, when L is
analyzed closer to the wall, it can be assumed that L=y. In addition, the destruction term
depends on the distance from the wall. This accounts for the high rate of dissipation in
nearness of solid boundaries. It is very important to note that the maximum dimension of
the dissipating eddies in the direction perpendicular to the flow must be equal to the
distance from the wall.
As mentioned earlier, the SA model for the destruction term is very similar to the
NK approach. The major difference in the derivation of the destruction term is the way
SA defines the non-dimensional function beside the constant in the term. The SA model
assumes that the blocking effect on the wall in the boundary layer is felt at a distance
through the pressure term. The pressure term acts as the main destruction term in the
Reynolds shear stress. For that reason the first term of the destruction term can be written
as; 1 ( / )2 , where c w1 is constant and d is the distance to the wall. To overcome

the problem with slow decay in the outer region, SA multiplied the destruction term by a

53

non-dimensional function f w which is equal to 1 in the log layer near the wall.
Consequently the new destruction term then becomes:
= 1 ( )2

(3-11)

Now that the production, diffusion, and destruction terms are defined, the rest of
the SA model will be explained. The relationship between all the working terms in the
equation and the turbulent kinematic eddy viscosity is = 1 =

function fv1 is defined as:

1 = 3 + 3

and the wall

(3-12)

The term is defined as the modified Vorticity magnitude that is maintained in

the buffer layer with log behavior. This is defined as:

2 2 2

, = 2 , = 2 = 1
(3-13)

1+1

The destruction term function fw is:


1

6
6
1+3

= 6 + 6
3

, = + 2 ( 6 ), = min

2 2

2 = 3 exp (4 2 )

, 10

(3-14)

(3-15)

Now that the SA model is completely defined as a one-equation model. SA


suggests the following constants to be used with its equation to do numerical simulation.
The suggested values for the constants are:

54

1 = 0.1355; 2 = 0.622; 1 =

1 1 + 2
+
; 2 = 0.3; 4 = 2 ; 1 = 7.1
2

3 = 1.2 ; 4 = 0.5 ; =

2
; = 0.41
3

3.3. Appling FLUENT Setup


All of the modeling and analysis was done using ANSYS FLUENT. Before all
this could be done the software had to be calibrated and initial parameters had to be
applied to the model within FLUENT. The following section will explain the setup,
application of boundary conditions to each model, and application of the mesh
refinement.

3.3.1. Initial Setup


In order to open FLUENT and start the CFD analysis, the user first has to open
ANSYS WORKBENCH as illustrated in Figure 2-27 and load the mesh from ANSYS
ICEM CFD. In order to do that, the user has to drag the correct cells in ANSYS
WORKBENCH to the workbench window. The cells are the geometry block and the
FLUENT block. The user must right click on the mesh cell to load the ANSYS ICEM
CFD mesh to the workbench. Then to load the mesh into FLUENT, the user must drag
the mesh cell to the FLUENT block and then double click on the setup cell to open
FLUENT. However, the FLUENT Launcher window will open first, and the user must
select the following options that are illustrated in Figure 3-1. Then press OK to start
FLUENT.

55

Figure 3-1:
Fluent Launcher option selection

As soon as FLUENT opens the user needs to set up the problem. Almost all the
steps are the same for each simulation except for the boundary condition setup that will
be discussed later. Figure 3-2 shows how to apply the problem setup within FLUENT to
get the best results.
The first thing that is done in the Problem Setup is the General setup. This is
where the mesh is checked; after that is completed the correct scale and the units are then
selected for the model. The reason why the correct scale and the units are selected is
because the model is in SI units and it needs to be converted to British units and scaled to
the correct size. In order to convert the units from SI to British units, the user must click
on the General tab and then select the Units menu. After that task is completed, the
user must scale the model to the correct size. In order to scale the model, the user must
click on the Scale menu and select ft for the View Length Unit In and then in the

56

scaling region the user must select the Convert Units option and units of inches. This is
illustrated in Figure 3-2. Then press Scale and close the dialog box by clicking Close.

Figure 3-2:
General setup step in Fluent

After the General setup is completed, the user must select the following steps to
complete Problem Setup. The steps are: Models, Materials, Cell Zone Conditions,
Boundary Conditions and Reference Values setups. The Model setup allows the user to
set various flow model options, e.g. phase change, mass transfer, etc. For this study, the
Viscous model is the only one selected. In the Viscous model option the SA turbulent
model is selected and the SA model uses the constants that are explained in Section 3.2.
This is illustrated in Figure 3-3.

57

Figure 3-3:
The Viscous Model Dialog box displaying the SA model setup

The next step is the Materials setup where the user must select the fluid and solid
materials. As mentioned earlier, the outside fluid (fluid box) for this thesis is going to be
air, and the solid will be set as aluminum in the Materials setup. The reason why
aluminum was set as the solid material, and not carbon fiber, is due to two reasons. First
FLUENT does not have carbon fiber in its data base. The second reason is because the
wall is assumed to be smooth and an arbitrary material is used.
The next steps in the Problem Setup are the Cell Zone Conditions and the
Reference Values setup. The Cell Zone Condition task allows setting the type of cell zone
condition parameters for each zone i.e. fluid domain is set as fluid. The Reference Value
Task page allows setting the reference quantities that are used for computing different
variables after the solution process has finished. Figure 3-4 illustrates the reference
values that are being used for the HPV model; however, the reference values for the other

58

models such as the oblate ellipsoid and the flat plate are all the same except for the area
and the velocity values that change with each simulation.

Figure 3-4:
Reference Values task for the HPV model

3.3.2. Boundary Conditions


The following discussion summarizes the Boundary Conditions task in FLUENT,
and Boundary Conditions for each simulation, are shown in Tables 3-1 and 3.2. Recall
that while creating the mesh in ICEM, the boundary types were then set for each face in
the domain. The right boundary plane (YZ plane in the positive X direction) is the inflow
of the flow field (
= ), and the left boundary plane is the outflow. The top and bottom

planes, and the XY plane in the positive Z direction are set as Symmetry planes, as well

59

the Symmetry plane for the fluid boundary. The exception is the simulation of the
moving ground plane as shown in Table 3-2, the bottom plane is defined as a wall for
these cases.
Symmetry boundary conditions are used when the physical geometry and the
expected pattern of flow solution have mirror symmetry in order to reduce the total
computational time and power needed for the simulation. In addition, symmetries are also
used to model zero-shear slip walls in viscous flow.
In the Problem Setup a Boundary Conditions Task can be opened and this where
the boundaries are specified for each region, this is done according to Tables 3-1 and 3-2
for the flat plate, oblate ellipsoid and the HPV fairing simulations.

plane Position

Name

Right (YZ plane @ X+)

Inflow

Type
Velocity
Inlet

Left (YZ plane @ X-)

Outflow

Outflow

Top (XZ plane @ Y+)

Top of the outer volume

Symmetry

Bottom (XZ plane @Y-)


Far side (XY plane @
Z+)
Symmetry (XY plane
@Z-)
Model

Bottom of the outer volume

Symmetry

Far side of outer volume

Symmetry

Symmetry
Model

Symmetry
Wall

Table 3-1:
Boundary type for the Flat Plate, Oblate Ellipsoid, and HPV fairing run without a
ground plane

60

plane Position

Name

Right (YZ plane @ X+)

Inflow

Type
Velocity
Inlet

Left (YZ plane @ X-)

Outflow

Outflow

Top (XZ plane @ Y+)

Top of the outer volume

Symmetry

Bottom (XZ plane @Y-)


Far side (XY plane @
Z+)
Symmetry (XY plane
@Z-)

Ground Plane

Wall

Far side of outer volume

Symmetry

Symmetry

Symmetry

Model

Model

Wall

Table 3-2:
Boundary type for the HPV fairing model with moving simulation ground plane

Figures 3-5 and 3-6 illustrate the velocity inlet and the outflow setup. This step is
performed on all the models. A Velocity Inlet boundary condition is used to define the
velocity and the scalar properties of the flow at the inlet. By clicking on Velocity Inlet
and setting the momentum parameter. In the momentum parameter the user must select
the following options; the Velocity Specification Method is set to the Magnitude,
Normal to Boundary, the Reference Frame setting is set as Absolute, and the
Velocity/Magnitude setting is set to the freestream velocity. As the Velocity Magnitude
varies the Modified Turbulent Viscosity varies with it. The following equations are used
to find the values that are illustrated in Table 3-3.

= 0.16

= 0.07
3

= ( )
2

61

1
8

(3-16)
(3-17)
(3-18)

In these equations I is the turbulence intensity, is the root-mean-square of the

1
2
2 2
turbulent (defined as; =
+ + , and is the mean flow velocity. The
3

turbulence intensity can be also found using the Reynolds number. The turbulence length
l is a physical quantity related to the size of the large eddies that contain the energy in
turbulent flow, and L is the length of the model. In order to find the modified turbulent
viscosity , the user must use Equations 3-16 and 3-17 to find I and l then plug these
values into Equation 3-18 to get the value for as illustrated in Table 3-3

u avg (f/sec)

l (ft)

Re

6.562E+00
1.312E+01
1.969E+01
3.281E+01
3.937E+01
4.593E+01
5.249E+01
5.867E+01
6.562E+01

3.268E-02
2.997E-02
2.849E-02
2.673E-02
2.612E-02
2.562E-02
2.520E-02
2.485E-02
2.451E-02

5.775E-01
5.775E-01
5.775E-01
5.775E-01
5.775E-01
5.775E-01
5.775E-01
5.775E-01
5.775E-01

3.301E+05
6.602E+05
9.903E+05
1.650E+06
1.981E+06
2.311E+06
2.641E+06
2.951E+06
3.301E+06

modified
turbulent
viscosity
[],(ft2/sec)
1.517E-01
2.782E-01
3.966E-01
6.202E-01
7.274E-01
8.325E-01
9.356E-01
1.031E+00
1.137E+00

modified
turbulent
viscosity
[],(m2/sec)
1.409E-02
2.584E-02
3.685E-02
5.761E-02
6.758E-02
7.734E-02
8.692E-02
9.581E-02
1.057E-01

Table 3-3:
Change in modified turbulent viscosity () with velocity.

62

Figure 3-5:
Boundary condition setup for the velocity inlet

The outflow boundary condition is used to define the flow that exits the region. At
this plane, the details of the pressure and the flow velocity are unknown before the
solution has been generated by FLUENT. The pressure outlet is set to outflow and the
flow rate weighting parameter is set to one; this specifies that 100% of the outflow is
leaving the bounded area. These steps are illustrated in Figure 3-6.

63

[38]

Figure 3-6:
Boundary condition setup for the outflow

3.3.3. Solution Setup and Mesh Adaption


The next step in the problem definition process is the Solution Setup. This setup is
the same for all the models. Once the Solution Setup is completed the user can proceed to
the mesh adaption process. Mesh adaption is performed on the model to ensure that the
mesh is sufficiently fine to obtain accurate form the simulation.
The first step in the Solution Setup is defining the Solution Method, and
specifying the various parameters associated with the calculation of the final result.
FLUENT provides implicit and explicit solver formlations from the drop down lists. The
Pressure-Velocity Coupling option contains the settings for this scheme, and the SIMPLE
algorithm (Semi-Implicit Method for the Pressure-Linked Equation) is used for all the
models. This algorithm is selected due to its relationship between velocity and pressure
corrections to help enforce mass conservation and to accurately obtain the pressure field.
In addition, the SIMPLE algorithm substitutes the flux correction equations into a

64

separate continuity equation in order to obtain a separate equation for the pressure
correction term in each cell.
The next step is selecting the Spatial Discretization of the convection terms in the
solution equation. Choices for these options were based on ME692 course notes. The SA
model has only four terms in the Spatial Discretization menu: Gradient, Pressure,
Momentum, and Modified Turbulent Viscosity term, in Table 3-4.

Spatial Discretization
Gradient
Pressure
Momentum
Modified Turbulent
Viscosity

Option
Green-Gauss Node-Based
Second Order
Second Order Upwind
Second Order Upwind

Table 3-4:
Spatial Discretization Setup

Following the Solution Method setup, the user must then apply the solution
controls. The user must set solution parameters for the under relaxation factors of
pressure, density, body forces, momentum, modified turbulent viscosity and turbulent
viscosity. Those factors were initially set at default values, and then if the residuals
continued to increase after four or five iterations, the under relaxation factors were then
reduced. The default under relaxation factors are shown in Table 3-5.

65

[38, 41]

Under Relaxation Factors


Pressure
0.3
Density
1
Body Forces
1
Momentum
0.7
Modified Turbulent Viscosity
0.8
Turbulent Viscosity
1
Table 3-5:
Default under Relaxation Factors

After all the preliminary steps are completed, the user then defines which flow
parameters will be monitored by FLUENT, and select where the solution will be
initialized from. The main parameters that need to be monitored are the residual, drag, lift
and moment solutions. The residual monitor allows us to set the absolute convergence
criteria for all the residuals that apply to the SA model. The residual monitor setup is
illustrated in Figure 3-7. The Convergence Absolute Criteria for all the residuals was set
to 1e-05. This was done to achieve the best results, the solution process continue until the
difference between successive values is less then 1e-05.
The drag, lift, and moment monitors all have the same setup. In order to setup the
drag monitor, the user must double click on the drag option and select the wall zone as
the geometry itself, i.e. HPV fairing or oblate ellipsoid. The user has to make sure that the
forces are monitored in the correct direction by setting the vector values. For these
models the vector value was 1 for X and 0 for Y and Z; this means that the forces are
monitored in the X direction or the flow direction.

66

Figure 3-7:
Setup of the Residual Monitors

To initialize the solution, the user must select where it will be computed from.
The program then references the frame and checks the initial values to see if they are
correct. In order to do that, the user has to follow a few steps. The first step is selecting
where the solution will be computed from, and what direction and zone the flow is
coming from and heading toward. This step is the same for all the simulations; the flow is
coming from the velocity inlet zone and going to the pressure outlet zone. For that reason
the user must select the Compute From as the Velocity Inlet zone. After the zone is
selected the initial values that are applicable variables to that zone are displayed. The
reference frame specifies whether the initial velocities are absolute velocities, by
selecting the Absolute option, or the velocity is relative to the motion of each cell zone
by selecting the Relative to Cell Zone option. For this study Relative to Cell Zone
option is used, due to the assumption that was made earlier that the model is stationary
and the air is moving around it at some speed. After the user has selected the correct
options, the user must then click Initialize to initialize the flow field.

67

In order to start the calculation and do mesh adaption, the user first has to start the
calculation by going to the Run Calculation task page and then checking the mesh to
make sure there no problems with the mesh. The user then needs to set the number of
iterations to 200 and click Calculate as illustrated in Figure 3-8. After these preliminary
results are produced they are used for the mesh adaption process. In order to do mesh
adaption, the user must click on the Adapt drop down menu, and the select the
Gradient Adaption option, as illustrated in Figure 3-9.
The gradient adaption option is a function that allows the user to mark specific
cells and adapt the mesh based on the gradient. For this thesis the method that is used to
adapt the mesh is gradient based with respect to velocity magnitude; all the adaptions
were done to refine the mesh, and to make sure the solution is grid independent. The
gradient method was selected because the momentum equation is the only governing
equation that is used for this simulation to find drag. Velocity gradient in the boundary
layer determine the drag on the surface, which is why velocity gradient were chosen to
control the adaption. Mesh adaption was done until the mesh reached independency and
the solution converged. That means the solution for drag was no longer dependant on the
mesh fineness. This process will be explained in depth in chapter 4.

68

Figure 3-8:
Run Calculation Setup

Figure 3-9:
Gradient Adaption / Mesh Adaption Setup

69

The mesh adaption process is illustrated in Figure 3-10, as the mesh changes from
the original one (A) to the mesh is used to set the final solution (D). The mesh adaption is
done every 200 iterations and data for viscous and pressure drag is then recorded for each
mesh as well.

Figure 3-10:
Mash adaption, A the original model, B D is the adaption every 200 iteration.

3.4. Solution to the Problem


The solutions to each model are provided in the results task page, where the user
can set up and display the results of the CFD simulation using FLUENT. The graphical
results that let the uses visually inspect the results are also generated using FLUENT.
Those graphical results include contours, vectors, pathlines, particle tracks, animations
and plots. In addition the user also has the ability to get the numerical solutions for the
drag and lift forces from the Reports task page.

70

3.4.1. Graphical and Numerical Solutions


The main kinds of graphical solution used for this thesis are the contours and the
vectors solutions. The vectors solutions let us inspect specific scalar fields as a vector
field or as a contour plot showing variations of velocity, vorticity, pressure, etc. In order
to find the drag and lift forces, the user must go to the report task page and select the
Forces option and then select the appropriate geometry for the wall zone and
appropriate vector directions to compute the forces. For the drag force the user must set
the vectors to X=1 and Y=Z=0, and for the lift force the user must set the vectors to Y=1
and X=Z=0.

3.5.Drag Calculation
The main aerodynamic force that is important in this thesis is the drag force.
Drag force is the net horizontal force with respect to direction of motion. The flow
domain is illustrated in Figure 3-11. This Figure shows domain and the coordinate system
around the HPV fairing; all other models use the same domain and coordinate system.
The fluid domain provides a basis for the calculation of the drag force by
FLUENT. The velocity inlet Y-Z plane (S 1 ) is assumed to be located in front of the
model, and far enough ahead of the model so that the incoming flow will satisfy the
undisturbed free stream conditions. The outflow Y-Z plane (S 2 ) is the downstream
condition, located sufficiently downstream of the model. All the walls of the virtual
tunnel form a uniform cross section that is parallel to the ongoing flow, and blowing and
suction is not allowed on any of these surfaces. Lastly, the flow in the domain is steady,
subsonic and incompressible. The fluid domain defined in FLUENT can be used to
71

describe a control volume which allow, the formulation of the drag force, as discussed
below.

S2
U
S1

Figure 3-11:
Control volume for derivation of drag equation using HPV model

The total drag on the body can be determined by calculating the change of
momentum equation in the direction of flow which in our case is the X-direction.

= ( + 2 ) ( + 2 )
1

(3-19)

Equation 3-19 represents the pressure forces driving the flow through the control
volume and the flux of momentum across the faces of the control volume. Where U is
the free stream velocity, u is the axial velocity, is the free stream density, local
density, and P is the static pressure. The conservation of mass for steady flow through
the control volume is given as:

72

1 = (
)

=0

(3-20)

The total pressure equation is given as:

= + ( 2 + 2 + 2 )

(3-21)

Where U, V and W are the velocity magnitude in the x, y and z direction. As


assumed earlier, flow at the S 1 plane corresponds to the undisturbed free steam
conditions. By substituting Equation (3-21) into Equation (3-19) , the total drag equation
can be obtained as and over all drag comes :

= ( + ) + (2 2 ) + ( 2 + 2 )
2
2
1
2
Equation 3-22 is based on Kusunose 1997 and FLUENT user guide

(3-22)

[44, 38]

, where

P is the total pressure from equation 15, P i is the free steam total pressure, and U is
the free stream velocity. If it is assumed that the upstream (velocity inlet) and the
downstream (pressure outflow) planes have the same area, the drag equation can be
rewritten (3-22) as:
= ( + ) + ( )
1

(3-23)

Equation 3-23 can be normalized by the dynamic pressure in order to produce the
non dimensional coefficient of drag.

= 0.5 2

73

(3-24)

C D is the coefficient of drag, FD is the drag force, q is the freestream dynamic


pressure, and S is the reference area. The non-dimensional quantity is a function of the
Reynolds number. The Reynolds number influences boundary layer behavior, especially
the transition to turbulence and the location of the separation point. Both of these factors
affect the total drag.
FLUENT provides the results of the total drag as a sum of two major components
where = + . The first component is the pressure drag coefficient, and the
second component is the viscous (or friction) drag.

[24]

For this thesis three major models i.e. flat plate, HPV fairing, and oblate ellipsoid
were used. For the flat plate the drag is totally due to the viscous drag.. For the oblate
ellipsoid and the HPV fairing the drag is composed of pressure and viscous contributions.
The reason why pressure drag plays such a major role is due to the flow separation
around the rear portion of these blunt objects.

74

Chapter 4: Baseline Solutions and Calibration of FLUENT

Before using and trusting CFD analysis techniques on a new configuration, one
should benchmark or validate the technique against a known test case similar to the new
configuration. If no appropriate test case exists, then the CFD analysis results would be
compared with another analysis or an experimental technique such as a wind tunnel. The
benchmark (baseline) test is a process of numerical analysis performed on a case which is
a replica of the previous results of numerical simulations or real time testing. In this
study, testing is performed by comparing test case results against published values in
Fluid Dynamics of Drag by Hoerner and standard fluids texts.
For CFD, the baseline process should result in guidelines for a specific class of
problems. The guidelines should describe what kind of turbulence model, preferred
boundary conditions, and meshing strategy (i.e. growth rate, cell size and clustering) are
required to achieve a desired level of precision and confidence in the results.
Following the procedure outlined in Chapter 3, results were generated for baseline
models, i.e. the oblate ellipsoid and the flat plate. The oblate ellipsoid and the flat plate of
very small thickness were modeled using SolidWorks CAD software to match the models
in Hoerner

[11]

in Chapter 1, Figure 1-2. Hoerner presented selective results on the drag of

rotationally symmetric bodies with different ratios of

based on wind tunnel tests.

Hoerners results are compared to the results of the numerical simulation using ANSYS
FLUENT. Flat plate results are compared to values found in standard fluid text.

75

4.1. FLUENT Calibration using Flat Plate

A CAD and mesh model of a flat plate with thickness of almost zero and the same
length as the HPV fairing was created. This flow case is simple but useful since the drag
is entirely due to wall shear stress, and this depends solely on the wall velocity gradient
in the boundary layer. The drag coefficient found for the flat plate was compared to the
drag coefficient found in Fundamentals of Fluid Mechanics by Munson.

[24]

This also

helped find the correct FLUENT parameters for the oblate ellipsoid such as initial setup,
materials for the fluid, geometry, boundary conditions, and dynamic mesh setup.
The type of mesh that was used for the flat plate is a combination of structured
and unstructured mesh. The structured mesh is used around the flat plate, i.e. for
boundary layer mesh, and the unstructured mesh is used everywhere else. The mesh was
created usingANSYS ICEM CFD, and this was also done for all other models such as the
oblate ellipsoid and the HPV fairing. The mesh for the flat plate is displayed in Figure 41.

76

Figure 4-1:
The computational mesh for the flat plate, overall view (A), and structured mesh
(boundary layer) (B)

The drag coefficient for the flat plate is found at various Reynolds numbers, and
then compared to the calculated drag for those Reynolds numbers based on correlation
analytical and empirical results. As noted above, the drag for the flat plate is composed of
100% viscous drag (and no pressure drag) because the flat plate had no flow separation
and thickness equal to zero. Four Reynolds numbers are selected to be compared, two
from the laminar region and two from the turbulent region. The laminar Reynolds
numbers are 1X105 and 2X105, and the turbulent Reynolds numbers are 1.5X106 and
2X106. Following the CFD process, the numerical values were then compared to the
calculated values to identify differences.

77

Table 4-1 shows the results for the drag coefficients found using FLUENT and
compared with the calculated drag and empirical drag. Table 4-2 compares the FLUENT
values with empirical values from Munson

[24]

. The calculated drag assumed to be fully

turbulent for the turbulent Reynolds number cases, i.e. there was no assumption made for
laminar to turbulent transition. The drag that was calculated used the following equations:

0.455

= (log

)2.58

1.328
0.5

, Laminar flow

(4-1)

, Turbulent flow, smooth plate

(4-2)

Percent error and percent difference were computed for each case using the following
equations:
% =

|()empirical|
empirical

100

% = 11 +22 100

(4-3)

(4-4)

Reynolds
number

CFD values
for C D

1.0E+05
2.0E+05
1.5E+06
2.0E+06

4.301E-03
3.036E-03
4.299E-03
4.140E-03

Calculated
values for C D
4.200E-03
2.969E-03
4.149E-03
3.940E-03

Percent error
for CFD vs.
calc
2.41%
2.23%
3.60%
5.07%

Percent Diff CFD


vs. calc
2.38%
2.20%
3.54%
4.95%

Table 4-1:
CFD drag data vs. calculated data for flat plate using integral method

78

Reynolds
number

CFD values
for C D

1.0E+05
2.0E+05
1.5E+06
2.0E+06

4.301E-03
3.036E-03
4.299E-03
4.140E-03

Empirical
values for C D
from Munson
4.400E-03
3.100E-03
4.200E-03
4.000E-03

Percent error
for CFD vs.
Munson
2.26%
2.08%
2.35%
3.51%

Percent diff CFD


vs. Munson
2.28%
2.10%
2.32%
3.45%

Table 4-2:
CFD drag data vs. Empirical data found from Fundamentals of Fluid Mechanics
5th edition by Munson for flat plate.

4.45E-03
4.30E-03
4.15E-03

coefficient of drag

4.00E-03
3.85E-03
3.70E-03

CFD Data
calc data
Munson Data

3.55E-03
3.40E-03
3.25E-03
3.10E-03
2.95E-03
1.0E+04

1.0E+05 2.0E+05
Reynolds number

1.0E+06 2.5E+06

Figure 4-2:
Comparison of three coefficients of drags for a flat plate

Figure 4-2 illustrates graphically what is displayed in Tables 4-1 and 4-2 for CFD
data that was found using ANSYS FLUENT, compared to the calculated and empirical
values.
79

Figure 4-3:
Empirical data for drag coefficient of a flat plate adopted from Fundamentals of Fluid
Mechanics 5th edition by Munson

Figure 4-3 illustrates the empirical data that is shown in Table 4-2. This figure
indicates how the drag coefficient behaves in turbulent flow with respect to surface
roughness, and Reynolds number.

4.2. FLUENT Calibration Using Oblate Ellipsoids

The oblate ellipsoid was used to calibrate ANSYS FLUENT and ANSYS ICEM
CFD, in order to find the best mesh density and ANSYS FLUENT parameters. Two
oblate ellipsoids geometries of

ratios 2 and 4 were modeled in SolidWorks to match the

tested solid geometries from Hoerners books and Figure 1-2. The mesh for each model
was generated, and the mesh sensitivity was found. That data from the oblate ellipsoid
was then used as a reference for the HPV fairing mesh.

80

The boundary condition types that were set in ANSYS ICEM CFD were also used
as the reference for the boundary condition setup for the HPV fairing. The Right
boundary (YZ plane @ X+) was set as the inflow of the flow field (
= ). The left

boundary plane is the outflow. The rest of the fluid boundaries, including the top plane,

bottom plane, the XY plane in the positive Z direction, and the Symmetry plane are set as
symmetry. The type of each boundary condition is specified in Table 3-1 in Chapter 3.
The boundary conditions that were created in ANSYS ICEM CFD in order to be
used in FLUENT are set as follows. Inflow boundary condition is set as the velocity inlet
to define the velocity and scalar properties of the flow at the inlet boundaries. Outflow
boundary condition is used to model the exit pressure and the flow velocity that is not
known prior to solution of the flow problem. The wall boundary condition is used to
model the actual physical geometry that is stationary, and the moving wall for ground
effect cases. Symmetry boundary conditions are used to model zero-shear slip walls in
the viscous flow. They are also used when the physical geometry and the expected flow
solution have mirror symmetry.
The following global mesh characteristics used for the oblate ellipsoid mesh
generation are: global element scale factor of 0.4 of an inch, and global element seed size
of 128 inches. Also during the Part Setup the oblate ellipsoid surface is the only surface
that will have a prismatic layer of maximum size of 3 inches, and 0.08-0.2 inch first cell
transverse dimension. The first cell transverse dimension depends on the total size of the
model, and the growth factor. For this thesis the growth factor is set to 1.08 for the
structured cells and 1.2 for the unstructured cells.

81

Figure 4-4:
The computational mesh for the oblate ellipsoids of ratio l/d=4 and 2

Figure 4-4 illustrates the computational mesh for the two oblate ellipsoids with
different

ratios, where a structured mesh is shown around the ellipsoid geometry and

the unstructured mesh is used everywhere else.

82

4.2.1 Results for the Oblate Ellipsoids

The flow field about the oblate ellipsoids ( ratios of 2 and 4) and the associated

drag forces are presented in this section. The Reynolds number was varied from 230
thousand to 3 million, and the freestream velocities were based on the Reynolds number,
and range from 1.98 to 59.64 ft/sec. These Reynolds numbers were selected because they
match the known values of drag that were done experimentally by Hoerner. The flow
field data are presented graphically in terms of velocity magnitude, pressure field, and
streamlines. The drag force is normalized by the freestream dynamic pressure and the
chord length of the ellipsoid and then compared with values in Hoerner.
The pressure field (dynamic and static) around both oblate ellipsoids is shown in
Figures 4-5 and 4-6 at Reynolds number 2 106 and V = 39.75 ft/sec. A low pressure

region (shown in blue) is present over and under the oblate ellipsoid. This is because the

angle of attack is 0 and it is symmetric about the X axis. There is also a stagnation
region as expected near the leading edge for both oblate ellipsoids that is indicated by
dark red in the static pressure plots. However, if one looks at the dynamic pressure plots
one can see the stagnation region is indicated by dark blue and the low dynamic pressure
regions are indicated by dark red. The velocity magnitude data for the same conditions as
Figure 4-5 are presented in Figure 4-7; the flow accelerates from the stagnation region
near the leading edge to the top of the oblate ellipsoid, and then the boundary layer is

visible on the rear portion of the body. It is clear that the fatter ellipsoid ( = 2) produces

a larger wake region.

83

Figure 4-5:
Pressure field (Static) about the oblate ellipsoids of ratios of l/d=2 and 4. Dark red
indicates stagnation regions and dark blue low pressure regions. Filled contour and
contour lines of oblate ellipsoid (A) Filled contour l/d=2; (B) contour lines l/d=2; (C)
Filled contour l/d=4 ;(D) contour lines l/d=4.
84

Figure 4-6:
Pressure field (Dynamic) about the oblate ellipsoids of ratios of l/d=2 and 4. Dark red
indicates low pressure regions and dark blue stagnation regions. Filled contour and
contour lines of oblate ellipsoid (A) Filled contour l/d=2; (B) Contour lines l/d=2; (C)
Filled contour l/d=4; (D) Contour lines l/d=4.

85

Figure 4-7:
Velocity magnitude about the oblate ellipsoids of ratios of l/d=2 and 4. Red indicates
high velocities and the light blue the stagnation region near the leading and trial edges.
Filled contour and contour lines of oblate ellipsoid (A) Filled contour l/d=2; (B)
Contour lines l/d=2; (C) Filled contour l/d=4; (D) Contour lines l/d=4.

86

However, when turbulent flow is analyzed in FLUENT, the program has a


problem in predicting the correct separation point, which then produces slightly higher
values for drag coefficient in the turbulent region compared to the experimental values.
This will be explained in more detail in Section 4.3. To verify how the flow behaves
around the ellipsoid, streamline plots are used as illustrated in Figure 4-8, which shows
no reverse flow on the surface. The boundary layer can be absorbed more clearly by
examining the vorticity contours as plotted in Figure 4-9. The boundary layer gets thicker
as it moves farther away from the leading edge of the ellipsoid. In addition, the boundary
layer at the trailing edge is about two to three times thicker than at the leading edge and
the flow is fully turbulent by that time.

Figure 4-8:
Streamlines about oblate ellipsoids (A) l/d=2; (B) l/d=4.
87

Figure 4-9:
Vorticity about oblate ellipsoids (A) l/d=2 and (B) l/d=4, showing the boundary layer
thickness on each model.

4.3. Comparison Between Hoerners data and CFD data

Tables 4-3, 4-4, 4-5, and 4-6 demonstrate how the optimum mesh was found for
the oblate ellipsoid cases and how each drag component changes after each level of mesh
adaptation. One can see that the viscous and pressure C D values change, however when
the values start to converge and the difference between them is less than 1% the mesh
adaptation process is then stopped. The final value is used as the CFD C D value to
compare with the empirical value for each Reynolds number and

88

value.

2.000E+06
l/d=2
Iterations C D pressure C D viscous
200
400
600
800

6.572E-03
6.411E-03
6.398E-03
6.388E-03

3.799E-03
3.028E-03
2.882E-03
2.881E-03

Total C D

mesh density (cells)

1.037E-02
9.439E-03
9.280E-03
9.268E-03

8.652E+05
1.350E+06
1.827E+06
2.308E+06

Table 4-3:
Pressure CD, Viscous CD and mesh density as separate entities for RE=2E6 for oblate
ellipsoids of fineness ratio 2

2.000E+06
l/d=4
Iterations C D pressure C D viscous
200
400
600
800

3.004E-03
2.830E-03
2.855E-03
2.844E-03

1.838E-03
1.773E-03
1.500E-03
1.463E-03

Total C D

mesh density (cells)

4.842E-03
4.602E-03
4.355E-03
4.308E-03

3.761E+05
8.532E+05
1.378E+06
1.835E+06

Table 4-4:
Pressure CD, Viscous CD and mesh density as separate entities for RE=2E6 for oblate
ellipsoids of fineness ratio 4

3.000E+06
l/d=2
Iterations C D pressure C D viscous
200
400
600
800

5.893E-03
5.892E-03
5.777E-03
5.761E-03

4.319E-03
3.898E-03
3.870E-03
3.861E-03

Total C D

mesh density (cells)

1.021E-02
9.790E-03
9.647E-03
9.622E-03

8.652E+05
1.356E+06
1.847E+06
2.319E+06

Table 4-5:
Pressure CD, Viscous CD and mesh density as separate entities for RE=3E6 for oblate
ellipsoids of fineness ratio 2

89

3.000E+06
l/d=4
Iterations C D pressure C D viscous
200
400
600
800

3.188E-03
3.175E-03
3.052E-03
3.048E-03

1.251E-03
1.235E-03
1.183E-03
1.182E-03

Total C D

mesh density (cells)

4.439E-03
4.410E-03
4.234E-03
4.230E-03

3.761E+05
8.426E+05
1.386E+06
1.899E+06

Table 4-6:
Pressure CD, Viscous CD and mesh density as separate entities for RE=3E6 for oblate
ellipsoids of fineness ratio 4

The oblate ellipsoids drag coefficient (C D ) values are shown in Figure 4-10 and
Table 4-7 as a function of Reynolds number for both

ratios and compared to the

experimental data from Hoerner. The agreement in the laminar range is within 3% and in
the turbulent range is within 8.5%. However, the CFD results for drag coefficient are
higher in the turbulent region. This is illustrated in Figure 4-10 and Table 4-7, for both

ratios.

90

9.90E-03

Coefficient of Drag based on surface area (CD)

9.30E-03
8.70E-03
8.10E-03
7.50E-03
l/d=4 Hoerner
6.90E-03

l/d=4 CFD
l/d=2 Hoerner

6.30E-03

l/d=2 CFD

5.70E-03
5.10E-03
4.50E-03
3.90E-03
1.00E+05

Reynolds number on a log10 scale

1.00E+06

Figure 4-10:
Drag Coefficient of the oblate ellipsoids vs. experimental data by Dr. Hoerner.

When looking at Figure 4-10; it is clear that there is a bigger deviation in the

= 2 plots. This is probably due to the greater importance of pressure drag to the overall

drag for lower

ratios. Therefore, the drag results found from FLUENT are more

sensitive to the prediction of the separation point in the back side of the body.
Table 4-7 tabulates the CFD and experimental C D values, including percent
differences. It should be noted that the Hoerner data was read from Figure 1-2. It was

91

hard to read the numbers and there is a small uncertainty of about 2-3% for the empirical
data that may add to the total percent difference.

Re #

l/d=2
CFD

2.5E+05
3.7E+05
5.0E+05
6.0E+05
2.0E+06
3.0E+06

CD
8.991E-03
8.380E-03
7.833E-03
7.687E-03
9.268E-03
9.622E-03

l/d=2
%
Hoerner Difference
for l/d=2
CD
8.900E-03
8.400E-03
8.000E-03
7.700E-03
8.500E-03
9.000E-03

1.01%
0.24%
2.12%
0.17%
8.65%
6.68%

Re #

l/d=4
CFD

CD
2.3E+05 4.708E-03
2.6E+05 4.361E-03
3.0E+05 4.077E-03
1.5E+06 4.639E-03
2.0E+06 4.308E-03
3.0E+06 4.230E-03

l/d=4
%
Hoerner Difference
for l/d=4
CD
4.800E-03
4.500E-03
4.100E-03
4.400E-03
4.100E-03
4.000E-03

1.94%
3.14%
0.56%
5.28%
4.95%
5.59%

Table 4-7:
CFD drag data for both oblate ellipsoids CFD and experimental data from Hoerner

92

Chapter 5: HPV Fairing Results

Following the procedures outlined in Chapter 3, and baseline solutions for the
oblate ellipsoids and the flat plate discussed in Chapter 4, FLUENT program setting to
obtain accurate drag values were established. This chapter provides a detailed overview
of the results for the HPV fairing simulations for various ground clearance and the
resulting drag and lift coefficients.
The results are presented with flow profiles, graphs, and tabulated values. The
discussion sheds light on the total drag on the HPV fairing at various ground clearances.
The plots and graphs for the HPV fairing simulation at = 3 106 and a ground

clearance of 15 inches are presented in detail in this chapter. Results for other ground
clearances can be found in Appendices B and C.

5.1. HPV CFD Test Results

The CFD simulation was conducted on the HPV fairing geometry with various
ground clearance ranging from 3 to 18 inches using the S-A turbulent model, for
these = 3 106 ( i.e. with inlet velocity of 40 mph, or 58.67 ft/sec). In addition to

these simulations, two benchmark tests were also conducted at ground clearances of 30

inches and 297 inches. This was done to find the drag coefficient corresponding to zero
ground effect. Since the process of CFD meshing is an iterative process, a number of
meshes were applied and tested. However, due to the limited time and computing power,
only two major benchmarking simulations were done (297 and 30 inches) to check for
mesh convergence.

93

5.1.1. HPV Fairing Benchmark Results

The benchmark CFD simulations for the HPV fairing were done for freestream
conditions (297 inches) and 30 inches away from the ground surface. This was done to
find the drag on the HPV fairing with zero or minimum ground interference. The data
could then be compared to the HPV fairing at different ground clearances corresponding
to realistic design condition.
The freestream ground clearance of 297 inches was selected because it was
assumed that at a distance of 3 away from the ground, there will not be
any ground interference on the fairing drag. In addition, another benchmark case was

selected at 30 inches away from the ground surface. This was estimated as the minimum
ground clearance where the ground interference would be negligible.
Figures 5-1 and 5-2 show the velocity vectors, and Figures 5-3 and 5-4 show the
velocity profiles for the two benchmark simulations with a freestream velocity of =

58.67

. The magnitude of the velocity vectors in

is indicated by the color scale, and

the length of each vector depends on the direction of the freestream velocity around the
HPV fairing. The velocity profile reveals that the highest velocities (in red) are
encountered in the bottom leading edge and on the upper surface as the flow accelerates
around the front portion of the fairing geometry. As the flow along the rear of the fairing
slows down a thicker boundary layer appears at the trailing edge. There is also a high
velocity region below the HPV fairing in both benchmark cases. At the trailing edge there
is a low velocity region that is indicated by a dark blue color, which indicates a
stagnation region. Essentially, this means that the fluid stops moving at the trailing edge
of the HPV fairing.
94

In the 30 inch benchmark case one can see the effects of ground proximity on the
velocity distribution underneath the HPV fairing. When looking at Figures 5-2 and 5-4
one can see how the velocity shape changes under the fairing, as the fairing boundary
layer starts to interact with the ground boundary layer. This interference would be
expected to affect the total drag on the fairing. When comparing the two benchmark
simulations (30 inches and freestream), a 2.6% difference is observed for the drag values,
and a 3.7% difference is seen for the lift values.

Figure 5-1:
Velocity Vectors about the HPV fairing in freestream

95

Figure 5-2:
Velocity Vectors about the HPV fairing with ground clearance of 30 inches

Figure 5-3:
Velocity magnitude of HPV fairing in freestream, red indicates high velocity
regions and dark blue indicates stagnation region at the trailing edge of the HPV
fairing.

96

Figure 5-4:
Velocity magnitude of HPV fairing with 30inches ground clearance, red
indicates high velocity regions and dark blue indicates stagnation region at the
trailing edge of the HPV fairing.

To confirm the accuracy of these drag values, mesh adaption was used as
discussed previously in Chapter 4. Tables 5-1 and 5-2 demonstrate how the optimum
mesh was found for each benchmark simulation by noting the change in each drag
coefficient component after each level of mesh adaption. The stopping criteria for the
simulation is when there is no longer a significant change (i.e. less than 1%) in between
the drag coefficients for each mesh. Also the major contributor to the total drag
coefficient is the pressure drag; for these cases the ratio of pressure drag to viscous drag
is roughly 5:1.

Ground Clearance 30 Inches


Iterations C D pressure C D viscous Total C D mesh density (cells)
6.506E-02 2.208E-02 8.714E-02
6.732E+05
200
7.004E-02 1.933E-02 8.937E-02
8.932E+05
400
7.463E-02 1.741E-02 9.204E-02
1.151E+06
600
7.858E-02
1.733E-02
9.592E-02
1.427E+06
800
7.636E-02 1.870E-02 9.506E-02
1.683E+06
1000
Table 5-1:
Drag coefficient components for HPV fairing with ground clearance of 30 inches

97

Ground Clearance 297


Iterations C D pressure C D viscous Total C D mesh density (cells)
6.389E-02 2.285E-02 8.673E-02
200
5.205E+05
7.294E-02 1.564E-02 8.857E-02
400
5.580E+05
7.334E-02
1.595E-02
8.929E-02
600
7.751E+05
7.563E-02 1.724E-02 9.287E-02
800
1.025E+06
7.413E-02 1.845E-02 9.258E-02
1000
1.311E+06
Table 5-2:
Drag coefficient components for HPV fairing at freestream

The static, dynamic, and total pressure fields around the HPV fairing for the
freestream case are displayed in Figures 5-5 to 5-7. Figure 5-5 shows the static pressure
contour around the HPV fairing at the symmetry plane. There is a high pressure region
that is indicated by the red color directly in front of the fairing; this represents the frontal
pressure, and it is caused by the air attempting to flow around the front of the fairing
nose. The low pressure regions are indicated in green and blue on top of the fairings
surface. This represents the flow that is accelerating above and below the fairing and it is
at a lower pressure compared to the front portion of the surface. This wake region plays a
major role in the total drag coefficient.

Figure 5-5:
Static pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 297 inches using contour lines(A) and filled contour (B). Red indicates
stagnation pressure and green-blue low pressure regions

98

Figure 5-6 shows the dynamic pressure contour; the low pressure region is
indicated using blue and green colors, and the high pressure regions are indicated using
red-orange. The low dynamic pressure regions are found in front of, behind, and on the
surface of the HPV fairing, i.e. in areas where the velocity is low .

Figure 5-6:
Dynamic pressure contours of the benchmark simulation at symmetry plane with
ground clearance of 297 inches using contour lines(A) and filled contour (B). The low
pressure region is indicated using blue and the high pressure regions are indicated using
red-orange

Figure 5-7 shows the total pressure contour; is indicated using red, the low
pressure wake region is indicated using yellow, light blue, and light green, and the low
pressure regions are indicated using color blue. As mentioned above, the low pressure
wake region is created when the boundary layer separates from the fairing. The size of
the wake region plays a very significant role on the total drag coefficient. If the size of
the wake region is reduced with increasing ground clearance, or the ground clearance
stays the same but aerodynamic enhancements are added, then the low pressure in the
wake region will then increase. This will cause the pressure differences on the fairing in
the flow direction to decrease, and this eventually helps reduce the pressure drag.

99

Figure 5-7:
Total pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 297 inches contour lines(A), and filled contour (B). The high pressure
around the HPV fairing on the surrounding air is indicated using red; the low pressure
wake region is indicated using yellow, light green and light blue; and low pressure is
indicated using blue.

The static, dynamic, and total pressure fields around the HPV fairing with a
ground clearance of 30 inches and at a freestream velocity of = 58.67

are

displayed in Figures 5-8 to 5-10. In those figures, one can see the pressure distribution
and ground effect on the fairing using frontal and isometric views. The pressure fields for
this case have a very similar pressure distribution as the freestream case. However, when
one examines the pressure field below the fairing, it appears to be influenced by the
ground surface boundary.

100

Figure 5-8:
Static pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 30 inches Isometric view of contour lines (A), filled contours (B), front
view of contour lines (C), filled contours (D).Red indicates stagnation pressure and
green-blue low pressure regions

Figure 5-9:
Dynamic pressure contours of the benchmark simulation at symmetry plane with
ground clearance of 30 inches. Isometric view of contour lines (A), filled contours (B),
front view of contour lines (C), filled contours (D).The low pressure region is indicated
using blue and the high pressure regions are indicated using red-orange

101

Figure 5-10:
Total pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 30 inches Isometric view of contour lines (A); filled contours (B); front
view of contour lines (C); filled contours (D).The high pressure around the HPV fairing
on the surrounding air is indicated using red, the low pressure wake region is indicated
using yellow and light green, and low pressure is indicated using blue.

The boundary layer formation on the HPV fairing can be observed from the
vorticity contours, shown in Figures 5-11 and 5-12 for the two benchmark simulations.
One can see how the boundary layer develops and thickens on the upper surface of the
HPV fairing for each benchmark simulation. In association with the velocity magnitudes
shown in Figure 5-3 and 5-4, the vorticity values represent the boundary layer as a
gradient of velocity.
Figure 5-11 shows the in freestream, case and Figure 5-12 shows the 30 inch case.
Comparing the two figures, the proximity of the ground appears to have a definite effect
on the size of the wake region downstream of the fairing.

102

Figure 5-11:
Vorticity contours of HPV fairing in freestream flow

Figure 5-12:
Vorticity contours of HPV fairing with ground clearance of 30 inches

The pressure coefficients for the top and bottom surfaces of the HPV fairing for
both benchmark simulations are shown in Figures 5-13 and 5-14. Note that the flow
direction in these figures is right to left.
Figure 5-13 show the pressure underneath the HPV fairing. When examining the
graph one can see that the ground effect is already present in the 30 inch benchmark,
there is a significant deviation between the two curves. In addition, a separation bubble is
present on the bottom of the HPV fairing that crates a negative value for the pressure

103

coefficient near the leading edge (point at 8.5 ft) before it then recovers to freestream
pressure toward the trailing edge of the fairing.
Figure 5-14 shows the pressure coefficient on the upper surface of the HPV
fairing. This pressure coefficient reaches a value nearly equal to 1 at the leading edge,
then it drops below 0 over the length of the fairing, and then starts recovering toward the
trailing. In addition, one can also see that the shape of the curves are nearly the same. It
may be that the ground clearance affects the location of the stagnation point on the nose
of the fairing, which then causes small changes down stream of this point.

0.1
0.0
-0.1

Cp (Coefficient of Pressure)

-0.2
-0.3
-0.4

Flow
Direction

-0.5
-0.6
-0.7
-0.8

freestream

-0.9

30 inches

-1.0
-0.1 0.4 0.9 1.4 1.9 2.4 2.9 3.4 3.9 4.4 4.9 5.4 5.9 6.4 6.9 7.4 7.9 8.4
Length of HPV fairing (ft)

Figure 5-13:
Pressure Coefficient the beneath the HPV fairing, 30 inch benchmark (blue) and
freestream benchmark (red)
104

1.0
0.9
0.8

freestream
30 inches

0.7

Flow
Direction

Cp (Coefficient of Pressure)

0.6
0.5
0.4
0.3
0.2
0.1
0.0
-0.1
-0.2
-0.3
-0.4
-0.5

-0.1 0.4 0.9 1.4 1.9 2.4 2.9 3.4 3.9 4.4 4.9 5.4 5.9 6.4 6.9 7.4 7.9 8.4
Length of HPV fairing (ft)

Figure 5-14:
Pressure Coefficient above the HPV fairing at 30 inch (blue), and freestream (red)

The computed skin (viscous) friction coefficient of the fairing for both benchmark
simulations is shown in Figure 5-15. The skin friction coefficient for the 30 inches
ground clearance simulation is clearly higher than the freestream case. This is a clear
indication of the ground effect, which results in a higher viscous drag coefficient (refer to
Tables 5-1 and 5-2). The highest values of the skin friction coefficient are found near the
leading edge of the fairing. This is because of the initial impact of the airflow on the
fairing and the development of the boundary layer. Following the initial impact, the skin
friction coefficient starts decreasing gradually over the majority of the fairing before it
spikes at the trailing edge.

105

0.01
30 inches

0.009

freestream

Cf (skin Friction Coefficient)

0.008

Flow
Direction

0.007
0.006
0.005
0.004
0.003
0.002
0.001

-3.64E-17
0

0.5

1.5

2.5

3.5

4.5

5.5

6.5

7.5

8.5

Length of HPV fairing (ft)

Figure 5-15:
Skin Friction Coefficient on the bottom of the HPV fairing at 30 inch (black), and
freestream (red)

5.1.2. HPV Fairing at Different Ground Proximities Results

The results of the HPV fairing with ground clearances ranging from 3 inches to 18
inches, at a constant freestream velocity of = 58.67

, are discussed in this section.

However, only the figures for the 15 inch ground clearance are shown in this section, and
detailed results for other ground clearances can be found in Appendix B.
Figures 5-16 and 5-17 show the velocity profile and the velocity vectors around
the fairing with 15 inch ground clearance. The velocity magnitude data reveals that the
highest velocities are on the top surface of the fairing and the leading edge below the
106

fairing as the flow accelerates. This is shown in red and dark orange. One can also notice
that the ground boundary layer interacts with the fairings boundary layer at the trailing
edge and this causes a reduction in velocity in this region. There is also a high velocity
region just after the leading edge of the fairing, and this is due to the boundary layer not
being fully developed yet. In addition, there is also a high velocity near the trailing region
above the fairing. This seems to be caused by the shape of the top surface of the fairing,
where the flow picks up some speed and then slows down again.
At the trailing edge of the fairing there is a low velocity region that is indicated by
light blue, which indicates a stagnation region. This means that the fluid stops moving
directly behind the HPV fairing, and a wake region is created in a very similar manner to
the benchmark simulations.

Figure 5-16:
Velocity Vectors about the HPV fairing with 15 inches ground clearance

107

Figure 5-17:
Velocity magnitude of HPV fairing with 15 inches ground clearance, red indicates high
velocity regions and blue indicates stagnation region at the trailing edge of the HPV
fairing.

Table 5-3 shows the pressure and viscous drag coefficient components of total
drag of the fairing with a ground clearance of 15 inches, at various points in the
simulation. When looking at Table 5-3 one can see the changes in viscous and pressure
drag coefficients at different mesh adaption levels. In order to stop the simulation, the
change between the drag components should be theoretically zero or very close to it. The
criteria is used as the stopping guideline for the simulation is the percent change between
the runs is less than 1%. Also it should be noted the major contributor to the total drag
coefficient is the pressure drag.

Iterations C D pressure
9.298E-02
200
9.294E-02
400
9.239E-02
600
9.165E-02
800
9.127E-02
1000

Ground Clearance 15
C D viscous Total C D mesh density (cells)
3.515E-02 1.281E-01
6.862E+05
3.686E-02 1.298E-01
9.221E+05
3.792E-02 1.303E-01
1.151E+06
3.926E-02 1.309E-01
1.377E+06
3.941E-02 1.307E-01
1.602E+06

Table 5-3:
Drag coefficient components for HPV fairing with ground clearance of 15 inches

108

The static, dynamic, and total pressure fields around the HPV fairing with ground
clearance of 15 inches and freestream velocity of = 58.67

are displayed in Figures

5-18 to 5-20. All pressure figures show the fairing in isometric and front view. The front
view shows how the pressure is distributed around the fairing, and the isometric view
shows the pressure distribution on the ground under the fairing.
Figure 5-18 shows the static pressure contours around the HPV fairing. There is a
high pressure region in front of the fairing, which is indicated by the red. It is caused by
the air attempting to flow around the front of the faring. As the air molecules approach
the front of the fairing they begin to compress, and in doing so raise the air pressure.
The low pressure regions are indicated with green and are located above and
below the HPV fairing. The main low pressure regions are located in the lower leading
edge and directly above the fairing, corresponding to the high velocity region discussed
previously. The pressure then starts to increase toward the trailing edge under the fairing.
This is because the pressure is trying to equalize around the fairing.

Figure 5-18:
Static pressure contours of 15 inches ground clearance simulation at symmetry plane.
Isometric view of contour lines (A); filled contours (B); front view of contour lines (C);
filled contours (D).Red indicates stagnation pressure (frontal pressure) and green low
pressure regions
109

Figure 5-19 shows the distrbution of dynamic pressure around the HPV fairing
with a ground cleanance of 15 inches. The low pressure regions are indicated using blue
and green and high presure regions are indicated using red and orange. Low dynamic
pressure regions are corrspond to the low velocity regions in the flow,and vice versa.

Figure 5-19:
Dynamic pressure contours of 15 inches ground clearance simulation at symmetry
plane. Isometric view of contour lines (A); filled contours (B); front view of contour
lines (C); filled contours (D).The low pressure region is indicated using blue and the
high pressure regions are indicated using red-orange

Figure 5-20 shows the total pressure around the HPV fairing with a ground
clearance of 15 inches. The low pressure regions are indicated using green, high presure
regions are indicated using red and orange, and the low pressure wake region is indicated
using light green and several shades of yellow.
The wake region plays a significant role in the total drag of the HPV fairing. The
size of the low pressure region behind the HPV fairing is directly related to the wake
region. If the wake region gets bigger the low pressure region will decrease. This is
because the ground proximity also decreased. This then causes the pressure difference of
the HPV fairing in the flow direction to increase, and total drag will also increase.

110

Figure 5-20:
Total pressure contours of 15 inches ground clearance simulation at symmetry plane.
Isometric view of contour lines (A); filled contours (B); front view of contour lines (C);
filled contours (D).The high pressure around the HPV fairing on the surrounding air is
indicated using red and orange; the low pressure wake region is indicated using yellow,
and light green; and low pressure is indicated using green.

The velocity contours on the top and bottom surfaces of the HPV fairing with a
ground clearance of 15 inches are shown are Figure 5-21. This can be associated with the
velocity magnitude found in Figure 5-17. Common knowledge dictates that vorticity is a
gradient of velocity. Thus the vorticity contours can be used to represent the boundary
layer around, above, and directly under, the fairing.
One can see how the boundary layer starts to develop on the ground directly under
the HPV fairing, as result of decreasing the height between the fairing and the ground.
The boundary layer of the HPV fairing has some interaction with the grounds boundary
layer, and the separating boundary core starts to disappear toward the rear of the fairing.
The separating boundary core disappears where the boundary layers of the fairing and the
ground are completely developed. The compression between all ground clearances is
found in Figure 5-22.
111

Figure 5-21:
Vorticity contours of HPV fairing with ground clearance of 15 inches

Figure 22:
Vorticity contours compression of HPV fairing, 3 inches (A), 6 inches (B), 9 inches (C),
12 inches (D), 15 inches (E), 18 inches (F), 30 inches (G), freestream (H)

5.1.3. Ground Clearance Effect on Pressure and Skin Friction


112

Figures 5-23 to 5-25 displayed below illustrate the comparison of the pressure
coefficient above and beneath the HPV fairing, and the skin friction coefficient data for
the fairing with different ground clearance ranging from 3 to 18 inches, and a constant
freestream velocity of = 58.67

Figures 5-23 and 5-24 show a comparison of the pressure coefficient beneath and

above the HPV fairing, as a function of ground clearance. The reduced ground clearance
causes reduced pressure beneath the fairing. In addition, the rear of the fairing also acts as
a diffuser; this is due to the low pressure region behind the fairing and the pressure starts
to rise at X<2.4 in Figure 5-23 coincide with the tapering of the cross-section. This assists
with the increased speed of the airflow between the ground and the HPV fairing. Pressure
drag is the major contributor to the total drag of the HPV fairing. There is also a
separation bubble beneath the fairing similar to the benchmark simulation of the 30 inch
and freestream cases. This creates a negative value for the pressure coefficient near the
leading edge, before the pressure then recovers to the freestream pressure towards the
trailing edge of the HPV fairing.

[13, 14]

Figure 5-24 shows that the pressure coefficient

behaves very similarly to the benchmark simulation at the top surface of the fairing, but
the curves are shifted as ground clearance changes. It is believed that the shift is caused
by the effect of the ground clearance on the location of the stagnation point on the nose of
the fairing.

113

Cp (Coefficient of Pressure)

0.1
0
-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
-0.7
-0.8
-0.9
-1
-1.1
-1.2
-1.3
-1.4
-1.5
-1.6
-1.7
-1.8
-1.9
-2

Flow
Direction
3 inches
6 inches
9 inches
12 inches
15 inches
18 inches

-0.1 0.4 0.9 1.4 1.9 2.4 2.9 3.4 3.9 4.4 4.9 5.4 5.9 6.4 6.9 7.4 7.9 8.4
Length of HPV fairing (ft)

Figure 5-23:
Pressure Coefficient beneath the HPV fairing with different ground clearance ranging
from 3 to 18 inches

114

1.0

3 inches

Cp (Coefficient of Pressure)

0.9
0.8

6 inches

0.7

9 inches

0.6

12 inches

0.5

15 inches

0.4

18 inches

Flow
Direction

0.3
0.2
0.1
0.0

-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
-0.7
-0.1 0.4 0.9 1.4 1.9 2.4 2.9 3.4 3.9 4.4 4.9 5.4 5.9 6.4 6.9 7.4 7.9 8.4
Length of HPV fairing (ft)

Figure 5-24:
Pressure Coefficient the above the HPV fairing with different ground clearance
ranging from 3 to 18 inches

Figure 5-24 shows a comparison of the skin friction coefficient on the center of
the underside of the HPV fairing at different ground clearances. As expected, the skin
friction coefficient increases as the ground clearance decreases. The skin friction
coefficient increases and the pressure decrease as a result of decreasing the distance
between the bottom of the HPV fairing and ground plane. The highest value of skin
friction coefficient is found in the leading edge, and this is due to the initial impact of the
airflow with the HPV fairing. Following the initial impact the skin friction coefficient
then gradually decreases before it spikes at the trailing edge as a function of ground
clearance.

115

0.0155
0.0145

18 inches

0.0135

15 inches

0.0125

12 inches
9 inches

0.0105

6 inches

0.0095

3 inches

Cf (skin Friction Coefficient)

0.0115

Flow
Direction

0.0085
0.0075
0.0065
0.0055
0.0045
0.0035
0.0025
0.0015
0.0005

-0.0005
0

0.5

1.5

2.5

3 3.5 4 4.5 5 5.5


Length of HPV fairing (ft)

6.5

7.5

8.5

Figure 5-25:
Skin Friction Coefficient beneath the HPV fairing with different ground clearance
ranging from 3 to 18 inches

116

Cp (Coefficient of Pressure)

0.1
0.0
-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
-0.7
-0.8
-0.9
-1.0
-1.1
-1.2
-1.3
-1.4
-1.5
-1.6
-1.7
-1.8
-1.9

Flow
Direction

3 inches
6 inches
9 inches
12 inches
15 inches
18 inches
30 inches
Freestream

-0.1 0.4 0.9 1.4 1.9 2.4 2.9 3.4 3.9 4.4 4.9 5.4 5.9 6.4 6.9 7.4 7.9 8.4
Length of HPV fairing (ft)

Figure 5-26:
Pressure Coefficient the beneath the HPV fairing for all the simulations

117

Cp (Coefficient of Pressure)

1.0
0.9

3 inches

0.8

6 inches

0.7

9 inches

0.6

12 inches

0.5
0.4
0.3
0.2

Flow
Direction

15 inches
18 inches
freestream
30 inches

0.1
0.0
-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
-0.7
-0.1 0.4 0.9 1.4 1.9 2.4 2.9 3.4 3.9 4.4 4.9 5.4 5.9 6.4 6.9 7.4 7.9 8.4
Length of HPV fairing (ft)

Figure 5-27:
Pressure Coefficient the above the HPV fairing for all the simulations

118

Cf (skin Friction Coefficient)

0.0155
0.0145

18 inches

0.0135

15 inches

0.0125

12 inches

0.0115

9 inches

Flow
Direction

6 inches

0.0105

3 inches

0.0095

30 inches

0.0085

freestream

0.0075
0.0065
0.0055
0.0045
0.0035
0.0025
0.0015
0.0005
-0.0005
0

0.5

1.5

2.5

3 3.5 4 4.5 5 5.5


Length of HPV fairing (ft)

6.5

7.5

8.5

Figure 5-28:
Skin Friction Coefficient for HPV fairing for all the simulations

5.1.4. Ground Clearance Effect on Drag and Lift

The drag and lift coefficients that result from the HPV fairing simulations of
different ground clearances are examined in the section in order to explain the results that
were achieved using ANSYS FLUENT. The following description of ground effect is
based on literature by J.B. Barlow.

[3]

The interaction of boundary layer between a

moving object and a ground surface depends strongly on the ground clearances.

119

B
A

Figure 5-29:
Effect of ground proximity (h) on the boundary layer of the body; (A) Large ground
clearance i.e. freestream; (B) Medium ground clearance i.e. normal road vehicles; (C) Small
ground clearance i.e. racecars; (D) Very small ground clearance when body boundary layer
starts to touch the ground . (Adapted from Low Speed Wind Tunnel Testing 3rd edition by
J.B. Barlow)

Figure 5-29 shows the effect of the ground on the body boundary layer for various
ground clearances. Figure 5-29A illustrates a case where the ground clearance is very
large and there is no interaction between the flow around the body and the ground. This
description would correspond to the benchmark simulations of freestream flow around
the fairing.
Figure 5-29B illustrates a case where the ground clearance is at medium height,
which is the case for a normal road vehicle. For this thesis HPV fairing ground clearance
ranges from 30 to 18 inches. Figure 5-29B shows the flow accelerating under the body,
then the flow starts to interact with the moving ground and a deformation occurs in the
bodys boundary layer. In addition to the deformation of the bodys boundary layer, there
is also a boundary layer generated along the moving ground. The speed at the edge of
ground boundary layer is not equal to zero, but rather it is equal to the velocity of the
body. This is because the local speed at the edge of the grounds boundary layer is equal

120

to the speed of the flow separated flow in the wake region behind the body. The shape of
the bodys boundary layer starts to change due to the presence of the grounds boundary
layer, with decreased ground clearance (h).
Figure 5-29C illustrates a case where the ground clearance is at small height,
which is the case for a racecar. For this thesis HPV fairing ground clearance ranges from
15 to 6 inches. Here one has a strong interaction between the bodys boundary and
grounds layer, due to the fact that there is almost no core separating them. For a smaller
ground clearance as illustrated in Figure 5-29D, the underbody will be hindered by the
combined ground and body boundary layer, and this will cause blockage and lower the
maximum speed beneath the underbody; this causes lower underbody downforce and
therefore effect of ground clearance on the max speed beneath the underbody increases
lift and drag on the fairing. This is observed in dynamic pressure plots in Figure 5-30.

121

Figure 5-30:
Dynamic Pressure contours compression of HPV fairing, 3 inches (A), 6 inches (B),
9 inches (C), 12 inches (D), 15 inches (E), 18 inches (F), 30 inches (G), freestream (H)

The computed lift and drag coefficients are plotted in Figure 5-31 for a freestream
velocity of = 58.67

versus all HPV fairing ground clearances normalized by the

fairing length. In addition, Table 5-4 shows the values for pressure, viscous, total drag
coefficients, while Table 5-5 shows the values for pressure, viscous and total lift
coefficients. The results for the force coefficients with various ground clearances follow
similar trends to the theoretical data found in Figure 1-6 in Chapter one. Significant drag
increases are seen as the normalized ground clearance (h/l) decreases below 0.3.

122

As expected, the lift and drag coefficients are the highest when the ground
clearance is the smallest, and have the highest percentage difference from the freestream
values. Having a streamlined HPV fairing shape this causes a blockage of air beneath the
fairing. This causes a lot of the flow near the ground to create a higher drag on the bottom
surface of the HPV fairing. However, due to the shape and the ground clearance, the flow
is no longer symmetric between the top and bottom surfaces. Because of that, there are
larger speeds and a low pressure region near the top and bottom surfaces of the HPV
fairing. This results in the positive lift being higher than the negative lift (downforce) and
higher drag.

1.50E-01

CD and CL vs normalized ground clearance

1.40E-01

Drag Coefficient

1.30E-01

Lift Coefficient

CD and CL

1.20E-01
1.10E-01
1.00E-01
9.00E-02
8.00E-02
7.00E-02
6.00E-02
0.03

0.30
Ground Clearness (h/L) on a Log10 scale

3.00

Figure 5-31:
Effect of ground proximity on the lift coefficient (CL), and drag coefficient (CD) of
HPV fairing. Where h is the ground clearance and L is the length of the HPV fairing
(99 inches).
123

Ground Clearance C D pressure C D viscous


297 (freestream)
7.413E-02
1.845E-02
30
7.636E-02
1.870E-02
18
9.027E-02
3.321E-02
15
9.127E-02
3.941E-02
12
9.385E-02
4.176E-02
9
9.443E-02
4.353E-02

Total C D % difference
9.258E-02
0.0%
9.506E-02
2.6%
1.235E-01
28.6%
1.307E-01
34.1%
1.356E-01
37.7%
1.380E-01
39.4%

9.786E-02

4.404E-02

1.419E-01

42.1%

9.985E-02

4.336E-02

1.432E-01

42.9%

Table 5-4:
Effect of ground proximity on the drag coefficient (CD) of HPV fairing data

Ground Clearance C L pressure


297 (freestream)
6.614E-02
30
6.853E-02
18
7.590E-02
15
8.838E-02
12
9.608E-02
9
1.104E-01
6
1.294E-01
3
1.414E-01

C L viscous
-9.048E-04
-8.436E-04
-4.862E-04
-4.070E-04
-4.170E-04
-3.293E-04
-3.082E-04
-2.963E-04

Total C L % difference
6.523E-02
0.0%
6.768E-02
3.7%
7.542E-02
14.5%
8.797E-02
29.7%
9.566E-02
37.8%
1.101E-01
51.1%
1.291E-01
65.7%
1.411E-01
73.5%

Table 5-5:
Effect of ground proximity on the lift coefficient (CL) of HPV fairing data

5.2. Estimation of Discretization Error

According to the American Institute of Aeronautics and Astronautics (AIAA)


guidelines for the definition of an error and uncertainty are as follows.
Uncertainty is defined as:
"A potential deficiency in any phase or activity of the modeling process that is
due to the lack of knowledge." (AIAA G-077-1998)
Error is defined as:
124

A recognizable deficiency in any phase or activity of modeling and simulation


that is not due to lack of knowledge. (AIAA G-077-1998)
There are three situations in this study that could possibly leave room for error.
The cases are as follows: 1. Computer round-off error, 2. Iterative convergence error, and
3. Spatial discretization error. The following section will discuss these possible errors and
how they could develop.
Computer round-off error: This is due to the accuracy at which numbers are
stored on a computer. However, with modern computers and its their ability to store data
with 32 or 64 bits, the round-off error is considered to be least significant compared to
the other errors.
Iterative convergence error: This occurs during the iterative process that is used
during the CFD simulation. Ultimately, it must have a stopping point. The error varies
with the solution variation at the end of the simulation. Tables 5-1, 5-2 and 5-3
demonstrate how stopping criteria is used to get the final values. The criteria that is used
as the stopping guideline for the simulation is the percent change between the runs is less
than 1%.
Spatial Discretization errors: These are also known as numerical errors. They
happen from representation of governing partial differential flow equations and other
physical models with algebraic expressions on a discrete spatial domain, also known as a
grid or mesh. As the mesh is refined the solution should no longer depend on the grid
size, and a solution should converge. This is called grid convergence, and it is very
important in finding levels of numerical errors in CFD solutions.

125

5.2.1. Discretization Error Calculation

According to the Fluids Engineering Division of ASME the following steps are
used to find the Discretization error.

[4]

Step one: Define the representative cell size h, where is the volume of the

cell, and N is the total number of cells used for this computation. The value of h is only
used for integral quantities such as drag or lift coefficients.
1

3
=
=1( )

( 5-1)

Step two: Select three significantly different sets of meshes that are demonstrated
by subscript 1, 2 and 3; this is done to find the values of main variables. Three sets of
meshes from all the HPV fairing simulations were selected; however, only the data for
the ground clearance of 15 inches will be shown here. The rest can be found in Appendix
C.
Step three: Let 1 < 2 < 3 and 21 = 2 /1 , 32 = 3 /2 , and then calculate

the apparent order, p, by using Equations 5-2, 5-3 and 5-4.


=

ln 21

ln

32
21
r

+ ()

() = ln r21

p
s
32

= 32
21

(5-2)
(5-3)
(5-4)

where 32 = 3 2 21 = 2 1 , and denotes the solution on the kth mesh.

Also it should be noted that q(p)=0 for the first guess for r=constant in Equation 5-2, and

126

this equation can be solved using fixed point iteration with initial guess equal to the first
term.
Step four: Calculate the extrapolated values from Equation 5-5. This also can be
32
done for
.

21

21 1 2

(5 -5)

21 1

Step five: Calculate the estimated errors using Equations 5-6 and 5-7, along with
the apparent order p relative approximate errors using Equation 5-8.
(12 )

21 =

(23 )

32 =

(5-6)

(5-7)

21 )
(
2 3

21

(5-8)

Step six: The fine grid convergence index is found by

21

1.2521

21 1

(5-9)

Table 5-6 illustrates this calculation procedure for three selected grids. As
mentioned previously the rest of the values are shown in Appendix C. The GCI value
represents the resolution level and how much the solution approaches the asymptotic
values.

[43]

Table 5-6 shows the successive grid refinement results and a reduced GCI less

than 5%. Therefore, it can be said that the solution on the finest grid resolution is nearly
grid independent.

127

15 inches
CD

CL

N 1 , N 2 , N 3 1601697, 1151324, 686151 1601697, 1151324, 686151


h 1 , h 2 ,h 3

0.33597, 0.37505, 0.44568

0.33597, 0.37505, 0.44568

r 21

1.1163

1.1163

r 32

1.1883

1.1883

f1

0.1309

0.0880

f2

0.1306

0.0888

f3
p

0.1281
12.3196

0.0946
10.4175

f ext 21

0.1310

0.0876

32

0.1310

0.0876

21

0.1933%

0.8960%

e a 32

1.9238%

6.5504%

e ext 21

0.0671%

0.4194%

2.1793%

7.9566%

0.0840%

0.5220%

f ext
ea

e ext 32
GCI fine

21

Table 5-6:
Calculation of Discretization error results for a ground clearance of 15 inches

5.3. Tradeoff Study Between Ground Clearances Drag and Stability for a Typical
HPV

This section will discuss what happens when the HPV fairing ground clearance
changes, and its effect on the drag, lift, and handling. Tables 5-4 and 5-5 illustrate the
percentage difference in drag and lift coefficients for the HPV fairing with various
ground clearances, compared to the freestream cases.

128

According to Race Cars Aerodynamics, by Katz, the handling of the vehicle


with similar body shape increases with a lower center of gravity and closer ground
proximity. However, the drag and lift coefficient increase as well.

[14, 15]

In order improve the stability of the HPV, it needs to be closer to the ground;
however, it will consequently have higher aerodynamic drag which limits top speed.
Ultimately, an HPV design has to balance these effects.
In order to estimate what is the best height for the HPV a few trade-off studies
were done. The first trade-off study estimates the power required as a function of velocity
for each ground clearance and compares it to the average human output of 221

used

by the 2010 HPV team. In order to find the required power the following equations are
used:
1

= = = 2 = 3
2

Here = 2.84 2 and = 2.37 3

(5-10)

are constants and C D is used from

the Table 5-4 for different ground clearances .required power values are calculated as a
function of velocity for each ground clearance, and compared with the average human
output, as shown in Figure 5-32.

129

Power as a function of velocity

250
225

Clearance=3"

200

Clearance=6"
Power [(lb-ft)/sec]

175

Clearance=9"
Clearance=12"

150

Clearance=15"

125

Clearance=18"
Clearance=30"

100

Clearance=free"
Average human power

75
50
25
0
0

10

20

30

40
50
Velocity (ft/sec)

60

70

Figure 5-32:
Theoretical power as a function of velocity for various ground clearances

130

80

245

Clearance=3"
Clearance=6"
Clearance=9"
Clearance=12"
Clearance=15"
Clearance=18"
Clearance=30"
Clearance=free"
Average human power

225

Power [(lb-ft)/sec]

205
185
165
145
125
105
85
65
65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

Velocity (ft/sec)

Figure 5-33:
Zoomed values from Figure 5-30 that shows the behavior of each line

The second trade-off study is to find the rollover speed for a corner with a 15ft
radius as a function of ground clearance. It is assumed that the ground clearance and
center of gravity location differ by a constant. Using an Excel spreadsheet for rollover
calculation provided by Dr. Robert Ryan and the CSUN HPV design team the rollover
speeds for ground clearances of 3, 6, 9, 12, 15, 18, and 30 inches were found and shown
in Figure 5-33 and Table 5-7.

131

Ground Clearance vs. rollover speed

31

3 inches

Rollover Velocity (ft/ssec)

29

6 inches
9 inches

27

12 inches
15 inches

25

18 inches
23

30 inches

21

19

17
3

12
15
18
21
Groound Clearance (inches)

24

27

30

Figure 5-34:
Rollover speed for various ground clearances of the HPV

Ground clearance to the


bottom of the HPV fairing
in inches

ground clearance and


seat height total in
inches

Rollover
velocity in
ft/sec

3
6
9
12
15
18
30

10
13
16
19
22
25
37

30.23
27.50
25.50
23.83
22.45
21.30
17.97

V in
ft/sec for
max
power
77
77.5
78
78.5
79.5
81
88.4

Table 5-7:
Rollover speed and maximum velocity to achieve average human power of
221.40 for various ground clearances of the HPV

132

Chapter 6: Conclusion
The flow fields around the HPV fairing at different ground proximities and oblate
ellipsoid baseline models have been computed using the ANSYS FLUENT CFD
package. This software solves the Reynolds Averaged NavierStokes equations for 3-D
flow with incompressible flow assumption using the one equation Spalart-Allmaras (SA)
turbulence model. The surfaces of the HPV fairing and the oblate ellipsoid were assumed
to be smooth. The computational domain was selected to be large enough, and the mesh
was fine enough, so that the flow calculated field is independent of these parameters,
which results in solution convergence.
The drag and lift coefficients for the HPV fairing at different ground clearances
and the drag coefficient of oblate ellipsoids were calculated from the CFD computations.
The computed values for the oblate ellipsoids were compared to experimental values
from Fluid Dynamics of Drag by Hoerner to verify that FLUENT was providing
correct values and that the mesh was sufficiently fine. Then the computed values for the
HPV fairing with different ground clearances ranging from 3 to 18 inches were compared
to the HPV fairing in freestream and 30 inches away from the ground.
Based on the results for the HPV fairing with different ground clearance, the
following can be concluded.
1. The streamlined shape of the HPV fairing does in fact help to reduce drag
coefficient; however, due to its shape when the ground proximity increases, a lot
of the flow is blocked near the ground and it creates a higher coefficient of drag is
created. Also, in addition to the increase in drag there is also an increase in lift
coefficient. This is because due to the ground proximity, the flow is no longer

133

symmetric and it will have an increased speed near the top of the HPV fairing and
low pressure region beneath the HPV fairing. This results in a more positive lift
coefficient.
2. The pressure coefficient (C P ) on the fairing bottom surface decreases as ground
clearance decreases. This is because as air tries to flow around the HPV fairing as
the ground becomes closer, less airflow is permitted to flow beneath the vehicle.
This then causes a low pressure region. Additionally, the pressure above the HPV
fairing also decreased due to the effect on velocity on this region.
3. As discussed in Chapter 5, the HPV fairings drag coefficient (C D ) and lift
coefficient (C L ) increase as ground clearance decreases. For example, the HPV
fairing with a ground clearance of 15 inches has a remarkable increase in C D from
0.0923 to 0.131. This value represents a nearly 34.13% increase in drag from the
benchmark simulation. In addition, the C L increases from 0.0652 to 0.0880. This
value is nearly 29.68 % higher than the C L in freestream.
4. To take advantage of the drag decrease afforded by greater ground clearance the
vehicle height is such that cornering stability is severely compromised. It may be
possible that changes to fairing geometry can be used as an alternative approach
to minimizing ground effect. This is recommended as a topic for further study.

134

References
1. Anderson, J.D. Jr., Computational Fluid Dynamics the Basics with Applications,
McGraw-Hill, Inc, 1st edition, February 1995.
2. Anderson, J.D. Jr., Fundamentals of Aerodynamics, McGraw-Hill, Inc, 3rd edition,
February 2001.
3. Barlow, J.B., Rae, W.H. Jr., and Pope, A., Low-Speed Wind Tunnel Testing, New
York, NY: John Wiley& Sons Inc, 3rd edition, 1999.
4. Celik, I.B., Procedure for Estimation and Reporting of Uncertainty Due to
Discretization in CFD Applications, ASME Journal of Fluids Engineering, Vol. 130,
July 2008.
5. Date, A.W., Introduction to Computational Fluid Dynamics, Cambridge University
Press, 2005.
6. DeMoss, J.A., Drag Measurements on an Ellipsoidal Body, Master Thesis, Virginia
Polytechnic Institute and State University, Blacksburg, VA, August 2007.
7. Diasinos, S., Barber, T.J., Leonardi, E., and Hall. S.D., A Two-Dimensional
Analysis of the Effect of a Rotating Cylinder on an Inverted Aerofoil in Ground
Effect, 15th Australasian Fluid Mechanics Conference, The University of Sydney,
Sydney, Australia, December 13-17, 2004.
8. Dress, D.A., Drag Measurements on a Laminar-Flow Body of Revolution in the 13inch Magnetic Suspension and Balance System, NASA Technical Paper 2895, 1989.
9. Fox, R.W., McDonald, A.T, and Pritchard, P.J., Introduction to Fluid Mechanics,
New York, NY: John Wiley& Sons Inc, 6th edition, July 2003.
10. Hamamoto, N., Nagayoshi, T., and Koike, M., Research on Aerodynamics Drag
Reduction by Vortex Generators, Mitsubishi Motors Technical Review, No.16:1116, January 2004.
11. Hoerner, S., Fluid Dynamic Drag, Published by the Author, 1965.
12. Hoerner, S., and Borst, H.V., Fluid Dynamic Lift, Mrs Liselotte A. Hoerner, 1985.
13. Hunco, WH., Aerodynamics of Road Vehicles, Warrendale, PA: SAE Int, 4th
edition, 1999.
14. Katz, J., Aerodynamics of Race Cars, Annual Reviews Fluid Mechanics Journal,
38:27-63, January 2006.
15. Katz, J., Race-Car Aerodynamics, Cambridge, MA, Bentley Publishing, 1995.
135

16. Fidkowski, F.K.., and Darmofal, L.D., Review of Output-Based Estimation and
Mesh Adaptation in Computation Fluid dynamics, AIAA Journal , Vol. 49, No. 4,
April 2011.
17. Krishnami, P.N., CFD Study of Drag reduction of a Generic Sport Utility Vehicle,
Master Thesis, California State University, Sacramento, CA, Fall 2009.
18. Little, R.P., Flight Simulator Database Population from Wind Tunnel and CFD
Analysis of a Homebuilt Aircraft, Master Thesis, California Polytechnic State
University, San Luis Obispo, CA, May 2006.
19. Mafi, M., Investigation of Turbulence Created by Formula One Cars with the Aid
of Numerical Fluid Dynamics and Optimization of Overtaking Potential ANSYS
Conference & 25th CADFEM Users Meeting 2007, Congress Center Dresden ,
Germany, November 21-23, 2007.
20. Milliken, W.F., and Milliken, D.L., Race Car Vehicle Dynamics, Warrendale, PA:
SAE Inc, August 1995.
21. Mohammadi, A., Computation of Flow Over a High Performance, Master Thesis,
California State University, Northridge, CA, May 2009.
22. Monsch, S.C., A Study of Induced Drag and Spanwise Lift Distribution for Threedimensional Inviscid Flow Over a Wing, Master Thesis, Clemson University, SC,
May 2007.
23. Munk, M.M., Fundamentals of Fluid Dynamics for Aircraft Designers, New York,
NY, The Ronald Press Company, 1929.
24. Munson, B.R., Young, D.F., and Okiishi T.H., Fundamentals of Fluid mechanics,
New York, NY: John Wiley& Sons Inc, 5th edition, 2006.
25. Panton,R.L., Incompressible Flow, New York, NY: John Wiley& Sons Inc, 3rd
edition, 2005.
26. Paparone, L., Tognaccini, R., A Method for Drag Decomposition from CFD
Calculations, ICAS 2002 Congress, pp 1113.1-1113.9, 2002.
27. Roy, C.J., Raju, A., and Hopkins, M.M., Estimation of Discretization Errors Using
the Method of Nearby Problems, AIAA Journal, Vol. 45, No. 6, June 2007.
28. Spalart, P.R., and Allmaras, S.R., A One-equation Turbulence Model for
Aerodynamic Flows, AIAA paper No.92-0439, January 1992.

136

29. Scibor-Rylski, A.J., Road Vehicle Aerodynamics, London, UK, Pentech Press,
1975.
30. Schlichting, H., Boundary-Layer Theory, McGraw-Hill, Inc, 7th edition, 1979.
31. Steenbergen, C.K., Vortices in favorable pressure gradients, Master Thesis, Delft
University of Technology, Delft, South Holland, Netherlands, July 2004.
32. Thompson, J.F., Soni, B.K.., and Weatherill, N.P., Handbook of Grid Generation,
CRC-Press, 1st edition, September 1998.
33. Versteeg, H.K.., and Malalasekera, W.,An Introduction to Computation Fluid
Dynamics The Finite Volume Method, Pearson Education Ltd, 2nd edition, February
2007.
34. Tutorial on CFD verification and Validation, NPARC Alliance CFD Verification
and Validation Web Site, 20 December 2010,
http://www.grc.nasa.gov/WWW/wind/valid/tutorial/tutorial.html
35. The Spalart Allmaras Turbulence Model, Langley Research Center, 15 April 2011,
http://turbmodels.larc.nasa.gov/spalart.html
36. History of Theoretical fluid Dynamics, Centrum Wiskunde & Informatica, 17
September 2010, https://www.cwi.nl/fluiddynamicshistory
37. ANSYS Tutorials: 12 August2011, https://www1.ansys.com/customer/default.asp
38. ANSYS FLUENT 12.0 User Guide, 2009.
39. ANSYS ICEM CFD 12.1 User Guide, 2009.
40. ANSYS WORKBENCH 12.1 User Guide, 2009.
41. California State University, Northridge, ME692 Course notes
42. Javaherchi, T., Review of Spalart-Allmaras Turbulence Model and its
Modifications, University of Washington, ME Department, March 2010.
43. Roache, P. J. , Quantification of Uncertainty in Computational Fluid Dynamics,
Annual Reviews Fluid Mechanics Journal, 29:123-60, 1997.
44. Kusunose, K.., Development of a Universal Wake Survey Data Analysis Code.
AIAA-1997-2294, pp. 617-626, 1997.

137

Appendix A
Boundary layer calculation using the integral approach:

The integral Equations:

Figure A-1:
The infinitesimal controlled volume of thickness dx for the boundary layer (adopted from
Schaums Outline of Fluid Mechanics )

Figure A illustrates an infinitesimal controlled volume of a thickness dx. The


continuity equation supplies the mass Flux top that crosses into the controlled volume
from the top.
top = out in =

0 u dy dx

(1)

The X-component momentum equation is:


=

Then it will become the following equation:

138

(2)

0 u2 dy dx U(x)

0 u dy dx

(3)

In the above equation the term pd and dpd are neglected because they are a
smaller order then the entire term because pd is very small since the assumed is to be
very small and d is then an order smaller ; also the term for the momentum =

() . In addition divide the entire equation (3) by (-dx) and the new equation
becomes the von Karman integral equation:
+

= ()

d
u
dx 0

dy

d 2
u
dx 0

dy

(4)

After using ordinary derivatives on equation (4), where the density is assumed to
be constant over the entire boundary layer and the is a function of x. as a result for a
flow over a flat plate with a zero pressure gradient , such as U(x)=U and
be simplified and put into the following form:
=

= 0 this can

0 ( )

(5)

For the velocity profile of u(x, y) for the specific flow, equation (5) along with
=

=0

lets both (x) and 0 (x) be determined. Where =y and u=0.99U

= 0 (1 ) = 0

= 0

(6)

(1

) = 0

(7)

d is the displacement thickness and it is the distance the streamline outside the
boundary layer is displaced due the slow moving fluid inside the boundary layer. is the
momentum thickness and it is the thickness of the fluid layer with the velocity U that

139

possesses the momentum lost because of the viscous effect. It is frequently used as the
characteristic length for the boundary layer.

= 2

(8)

Laminar and Turbulent Boundary Layer:


The main boundary conditions that need to be meet for the velocity profile in the

laminar boundary layer for the flat plate with a zero pressure gradient are:
u=0

at

y=0

u=0.99U

at

y=

at

y=

=0

As figure B illustrates a general flow over a flat plate with uniform velocity and
the development of the boundary layers along the flat plate.

Figure A-2:
Boundary layer development along a flat plate. (Adopted from Fluid Mechanics
for Engineering a Graduate Textbook by Meinhard T. Schobeiri)

Laminar boundary layers:


140

According Prandtl/Blasius boundary layer solution that can be solved for by the
governing Navier-Stokes equations with negligible gravitational effects. They become the
following two equations:

+ (

+ (

2
2

2
2

(9)
(10)

Figure A-3:
Typical characteristics of boundary layer thickness () and wall shear wall stress
(w) for laminar and turbulent boundary layer. (Adopted from Fundamentals of
Fluid Mechanics 5th edition by Munson)

Using conservation of mass equation for incompressible flow becomes.


In addition Where v<<u

=0

and

(11)
(12)

141

Although both the boundary equations (11) and (12) and Navier-Stokes equations
(9) and (10) are non-linear partial differential equations, there a considerable difference
between them. For one, the y momentum equation has been eliminated and only leaves
the x momentum equation. The pressure variable has been eliminated and only leaving
the x and y velocity components as the only unknowns. In addition for the boundary layer
flow over the flat plate the pressure is assumed to be constant and the flow represents a
balance between viscous and inertial effects with the pressure playing no role.
For the laminar boundary layer, a parabolic velocity profile is assumed due to the
fact that the boundary layer is very thin.

= + + 2

(13)

The above boundary conditions are used to find the values for A, B and C
0=A
1=A+B+C2
0=B+2C
The solution of the problem then:
=0

This results in the laminar flow velocity profile:

1
2

= 2 2

(14)

Then substitute equation (14) in to equation (5) and the result of it is:
= 2

0 2 2 1 2 2 = 15 2

142

(15)

The wall shear stress is given by the following expression:


=

=0

(16)

Then equate the equation (15) and (16) to obtain


=

15

(17)

Then integrate equation (17) with =0 at x=0 to find the expression for (x) in the

laminar boundary layer, where = :

( ) = 5.48

(18)

To find the shear stress ( w(x)), local skin friction coefficient (c f ) and
dimensionless drag force that is the skin friction coefficient (C f ) substitute equation (18)
into equation (16).
( ) = 0.3652

() = .
= .

0.3652

(19)

(20)

= .
= . = .

(21)

Buffer boundary Layer:


Buffer boundary layer is the layer in between the laminar and turbulent boundary

layers; it is sometimes call the transition layer this is shown in figure B. The boundary
layer thickness increases in proportion to where x denotes the distance from the
leading edge according to Boundary-Layer theory by Schlichting. Near the leading edge

of the flat plate the flow is always laminar and it is becoming turbulent further

143

downstream on the plate. The transition takes place at distances x from the leading edge
and it determined by:
, =

And |= determined by:


= .

(22)

(23)

In addition the process of transition also involves a large decrease in shape factor
H 12 = 1 / 2 where 2 = and 2 = d

Figure A-4:
Changes in the shape factor (adopted from Boundary-Layer Theory 7th edition by H.
Schlichting)

Using Prandtl-Schlichting local skin friction equation for a Smooth Plates that is
given by:
=

( ).

(24)

Where the A is the function of only the transition Reynolds number and it is given
by:
=

144

(25)

Where local shear stress is a function of the local skin friction and is given by:

= ; =

(26)

Turbulent boundary layer:


In the turbulent boundary layer it is often assume a power law velocity profile for

this study the maximum R e number was found to be 2.95x106 so the n=7

(27)

Then substitute this velocity profile equation (27) into equation (5) and integrate
to get the following expression for w :
=

(28)

This w from the velocity profiles yields a =

= = so this

expression cannot be used at the wall. A new expression is needed to be derived for w ;
the Blasius formula is selected for local skin friction coefficient for turbulent model and
given by:
= .

(29)

In addition to the local skin friction coefficient an experimentally determined


formula for shear stress is given by:
=

(30)

= 0.231

(31)

Combine equation (28) with equation (30) and we get:


1
4

145

1
4

Equation

(31)

1
0 4

4
0 0.231 ;

can

be

integrated

=0

from

= 0.37

1
5

and

x=0

< 107

to

obtain
(32)

Equation (29) can substitute into equation (32) to get the local skin friction
coefficient to be:
() = 0.0577

1
5

< 107

(33)

Equation (33) can be substituted into equation 30 and divided by 2 and right side
multiplied by 2 to get the shear stress
() =

0.028852

1
5

< 107

(34)

In addition the dimensionless drag force that is the skin friction coefficient (C f )
becomes:
() = 0.072

1
5

< 107

(35)

Law of the Wall:


Law of the wall is the average velocity of the flow at a specific point. It is

proportional to the logarithm of the distance a point to the wall (flat plate surface). The
law was first published in the early 1930s by von Karman; it only applies to the flow that
is close to the wall. The general logarithmic equations are:
+ =

Where:

+ =

146

(36)

y+ is the wall coordinate, the dimensionless distance y to the wall


u+ is the dimensionless velocity
w is the wall shear stress
is the fluid density
u t is the friction velocity
For laminar study the y+ is assumed to be 5, for the buffer layer the y+ =11 and,
turbulent the y+=30

147

Appendix B
In Appendix B we can find all of the figures for velocity, pressure and vectors
contours from ANSYS FLUENT for HPV fairing at ground clearance of 3,6,9,12,18
inches.

Figure B-1:
Velocity Vectors about the HPV fairing with ground clearance of 3 inches

Figure B-2:
Velocity contours about the HPV fairing with ground clearance of 3 inches

148

Figure B-3:
Static pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 3 inches Isometric view of contour lines (A); filled contours (B); front view
of contour lines (C); filled contours (D).

Figure B-4:
Dynamic pressure contours of the benchmark simulation at symmetry plane with
ground clearance of 3 inches. Isometric view of contour lines (A); filled contours (B);
front view of contour lines (C); filled contours (D).

149

Figure B-5:
Total pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 3 inches Isometric view of contour lines (A); filled contours (B); front view
of contour lines (C); filled contours (D).

Figure B-6:
Vorticity contours of HPV fairing with ground clearance of 3 inches

150

Figure B-7:
Velocity Vectors about the HPV fairing with ground clearance of 6 inches

Figure B-8:
Velocity contours about the HPV fairing with ground clearance of 6 inches

151

Figure B-9:
Static pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 6 inches Isometric view of contour lines (A); filled contours (B); front view
of contour lines (C); filled contours (D).

Figure B-10:
Dynamic pressure contours of the benchmark simulation at symmetry plane with
ground clearance of 6 inches. Isometric view of contour lines (A); filled contours (B);
front view of contour lines (C); filled contours (D).

152

Figure B-11:
Total pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 6 inches Isometric view of contour lines (A); filled contours (B); front view
of contour lines (C); filled contours (D).

Figure B-12:
Vorticity contours of HPV fairing with ground clearance of 6 inches

153

Figure B-13:
Velocity Vectors about the HPV fairing with ground clearance of 9 inches

Figure B-14:
Velocity contours about the HPV fairing with ground clearance of 9 inches

154

Figure B-15:
Static pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 9 inches Isometric view of contour lines (A); filled contours (B); front view
of contour lines (C); filled contours (D).

Figure B-16:
Dynamic pressure contours of the benchmark simulation at symmetry plane with
ground clearance of 9 inches. Isometric view of contour lines (A); filled contours (B);
front view of contour lines (C); filled contours (D).

155

Figure B-17:
Total pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 9 inches Isometric view of contour lines (A); filled contours (B); front view
of contour lines (C); filled contours (D).

Figure B-18:
Vorticity contours of HPV fairing with ground clearance of 9 inches

156

Figure B-19:
Velocity Vectors about the HPV fairing with ground clearance of 12 inches

Figure B-20:
Velocity contours about the HPV fairing with ground clearance of 12 inches

157

Figure B-21:
Static pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 12 inches Isometric view of contour lines (A); filled contours (B); front view
of contour lines (C); filled contours (D).

Figure B-22:
Dynamic pressure contours of the benchmark simulation at symmetry plane with
ground clearance of 12 inches. Isometric view of contour lines (A); filled contours (B);
front view of contour lines (C); filled contours (D).

158

Figure B-23:
Total pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 12 inches Isometric view of contour lines (A); filled contours (B); front view
of contour lines (C); filled contours (D).

Figure B-24:
Vorticity contours of HPV fairing with ground clearance of 12 inches

159

Figure B-25:
Velocity Vectors about the HPV fairing with ground clearance of 18 inches

Figure B-26:
Velocity contours about the HPV fairing with ground clearance of 18 inches

160

Figure B-27:
Static pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 18 inches Isometric view of contour lines (A); filled contours (B); front view
of contour lines (C); filled contours (D).

Figure B-28:
Dynamic pressure contours of the benchmark simulation at symmetry plane with
ground clearance of 18 inches. Isometric view of contour lines (A); filled contours (B);
front view of contour lines (C); filled contours (D).

161

Figure B-29:
Total pressure contours of the benchmark simulation at symmetry plane with ground
clearance of 18 inches Isometric view of contour lines (A); filled contours (B); front view
of contour lines (C); filled contours (D).

Figure B-30:
Vorticity contours of HPV fairing with ground clearance of 18 inches

162

Appendix C

2.500E+05
Iterations
200
400
600
800

C D pressure C D viscous
7.219E-03
5.821E-03
5.814E-03
5.814E-03

2.889E-03
3.248E-03
3.225E-03
3.238E-03

Total C D

mesh density (cells)

1.011E-02
9.070E-03
9.039E-03
8.991E-03

8.652E+05
1.351E+06
1.788E+06
2.244E+06

2.600E+05
Iterations
200
400
600
800

C D pressure C D viscous
2.172E-03
2.083E-03
2.057E-03
2.024E-03

2.340E-03
2.340E-03
2.338E-03
2.336E-03

Total C D

mesh density (cells)

4.512E-03
4.423E-03
4.395E-03
4.361E-03

3.761E+05
8.124E+06
1.395E+06
1.842E+06

3.000E+05
Iterations
200
400
600
800

C D pressure C D viscous
2.145E-03
1.902E-03
1.868E-03
1.851E-03

2.231E-03
2.229E-03
2.227E-03
2.226E-03

Total C D

mesh density (cells)

4.376E-03
4.131E-03
4.095E-03
4.077E-03

3.761E+05
7.851E+05
1.362E+06
1.793E+06

1.500E+06
Iterations
200
400
600
800

C D pressure C D viscous
3.734E-03
3.669E-03
3.664E-03
3.647E-03

1.250E-03
1.094E-03
1.024E-03
9.922E-04

Total C D

mesh density (cells)

4.985E-03
4.763E-03
4.688E-03
4.639E-03

3.761E+05
8.037E+05
1.310E+06
1.778E+06

Tables C-1:
Pressure CD and Viscous CD as separate entities for oblate ellipsoids of finesse ratio 4

163

2.500E+05
Iterations
200
400
600
800

C D pressure C D viscous
7.219E-03
5.821E-03
5.814E-03
5.814E-03

2.889E-03
3.248E-03
3.225E-03
3.238E-03

Total C D

mesh density (cells)

1.011E-02
9.070E-03
9.039E-03
8.991E-03

8.652E+05
1.351E+06
1.788E+06
2.244E+06

3.700E+05
Iterations
200
400
600
800

C D pressure C D viscous
4.701E-03
4.418E-03
4.433E-03
4.421E-03

2.752E-03
2.761E-03
2.702E-03
2.705E-03

Total C D

mesh density (cells)

7.454E-03
7.179E-03
7.135E-03
7.126E-03

8.652E+05
1.344E+06
1.835E+06
2.291E+06

5.000E+05
Iterations
200
400
600
800

Iterations
200
400
600
800

C D pressure C D viscous
5.844E-03
3.876E-03
3.835E-03
3.835E-03

1.989E-03
2.421E-03
2.374E-03
2.372E-03

Total C D

mesh density (cells)

7.833E-03
6.297E-03
6.209E-03
6.207E-03

8.652E+05
1.341E+06
1.811E+06
2.248E+06

6.000E+05
C D pressure C D viscous Total C D mesh density (cells)
5.654E-03 1.883E-03 7.537E-03
8.652E+05
5.655E-03 2.209E-03 7.864E-03
1.337E+06
5.500E-03 2.210E-03 7.710E-03
1.826E+06
5.477E-03 2.210E-03 7.687E-03
2.291E+06

Tables C-2:
Pressure CD and Viscous CD as separate entities for oblate ellipsoids of finesse ratio 2

164

Iterations
200
400
600
800
1000

Iterations
200
400
600
800
1000

Iterations
200
400
600
800
1000

Iterations
200
400
600
800
1000

Iterations
200
400
600
800
1000

Ground Clearance 3
C D pressure C D viscous Total C D mesh density (cells)
9.577E-02
3.720E-02 1.330E-01
4.204E+05
9.829E-02
4.114E-02 1.394E-01
6.415E+05
9.749E-02
4.226E-02 1.398E-01
8.714E+05
9.818E-02
4.336E-02 1.415E-01
1.092E+06
9.985E-02
4.336E-02 1.432E-01
1.581E+06

C D pressure
9.318E-02
9.942E-02
9.825E-02
9.635E-02
9.786E-02

Ground Clearance 6
C D viscous Total C D mesh density (cells)
3.540E-02 1.286E-01
7.308E+05
3.544E-02 1.349E-01
9.420E+05
4.036E-02 1.386E-01
1.167E+06
4.426E-02 1.406E-01
1.384E+06
4.404E-02 1.419E-01
1.603E+06

C D pressure
9.229E-02
9.444E-02
9.460E-02
9.461E-02
9.443E-02

Ground Clearance 9
C D viscous Total C D mesh density (cells)
3.585E-02 1.281E-01
9.459E+05
3.782E-02 1.323E-01
1.177E+06
4.112E-02 1.357E-01
1.405E+06
4.305E-02 1.377E-01
1.634E+06
4.353E-02 1.380E-01
1.858E+06

Ground Clearance 12
C D pressure C D viscous Total C D mesh density (cells)
8.901E-02
3.891E-02 1.279E-01
8.986E+05
9.392E-02
3.578E-02 1.297E-01
1.154E+06
9.420E-02
3.979E-02 1.340E-01
1.368E+06
9.404E-02
4.146E-02 1.355E-01
1.598E+06
9.385E-02
4.176E-02 1.356E-01
1.821E+06
Ground Clearance 18
C D pressure C D viscous Total C D
9.318E-02
9.377E-02
9.265E-02
9.105E-02
9.027E-02

2.714E-02
2.741E-02
3.022E-02
3.287E-02
3.321E-02

1.203E-01
1.212E-01
1.229E-01
1.239E-01
1.235E-01

mesh density (cells)


6.890E+05
9.073E+05
1.127E+06
1.343E+06
1.581E+06

Tables C-3:
Pressure CD and Viscous CD as separate entities for HPV fairing
165

3 inches
N1 , N2, N3
h 1 , h 2 ,h 3

CD

CL

1580631, 871408, 530392

1580631, 871408, 530392

0.33312, 0.40626, 0.51799 0.33312, 0.40626, 0.51799

r 21

1.2196

1.2196

r 32

1.2750

1.2750

f1

0.1432

0.1411

f2

0.1398

0.1427

f3
p

0.1330
2.1284

0.1515
6.5182

f ext 21

0.1498

0.1405

f ext 32

0.1498

0.1405

e a 21

2.4097%

1.1672%

e a 32

4.8531%

6.1572%

4.3828%

0.4428%

11.2155%

7.8713%

5.7297%

0.5510%

e ext

21

e ext

32

GCI fine

21

Table C-4:
Calculation of Discretization error for HPV fairing with ground clearance of 3 inches

166

6 inches
CD

CL

N 1 , N 2 , N 3 1603025, 1167402, 730782 1603025, 1167402, 730782


h 1 , h 2 ,h 3

0.33293, 0.37005, 0.39815

0.33293, 0.37005, 0.39815

r 21

1.1115

1.1115

r 32

1.1690

1.1690

f1

0.1419

0.1291

f2

0.1386

0.1302

f3
p

0.1286
5.4358

0.1417
13.6966

f ext 21

0.1461

0.1287

f ext 32

0.1461

0.1287

e a 21

2.3113%

0.9114%

e a 32

7.2359%

8.8374%

2.8911%

0.2810%

11.9999%

10.1380%

3.7215%

0.3503%

e ext

21

e ext

32

GCI fine

21

Table C-5:
Calculation of Discretization error for HPV fairing with ground clearance of 6 inches

167

9 inches
CD

CL

N 1 , N 2 , N 3 1858015, 1405243, 945865 1858015, 1405243, 945865


h 1 , h 2 ,h 3

0.31791, 0.34892, 0.39815

0.31791, 0.34892, 0.39815

r 21

1.0976

1.0976

r 32

1.1411

1.1411

f1

0.1380

0.1101

f2

0.1357

0.1103

f3
p

0.1281
7.5874

0.1183
26.1866

f ext 21

0.1401

0.1100

f ext 32

0.1401

0.1100

e a 21

1.6187%

0.2160%

e a 32

5.5915%

7.2433%

1.5522%

0.0207%

8.5614%

7.4972%

1.9709%

0.0258%

e ext

21

e ext

32

GCI fine

21

Table C-6:
Calculation of Discretization error for HPV fairing with ground clearance of 9 inches

168

12 inches
N1 , N2, N3
h 1 , h 2 ,h 3

CD

CL

1820694, 1368402, 898603

1820694, 1368402, 898603

0.32101, 0.35307, 0.406199 0.32101, 0.35307, 0.406199

r 21

1.0999

1.0999

r 32

1.1505

1.1505

f1

0.1356

0.0957

f2

0.1340

0.0957

f3
p

0.1279
7.6712

0.1011
29.6691

f ext 21

0.1371

0.0957

f ext 32

0.1371

0.0957

e a 21

1.2012%

0.0835%

e a 32

4.5308%

5.5871%

1.1045%

0.0053%

6.7195%

5.6809%

1.3961%

0.0066%

e ext

21

e ext

32

GCI fine

21

Table C-7:
Calculation of Discretization error for HPV fairing with ground clearance of 12 inches

169

18 inches
CD

CL

N 1 , N 2 , N 3 1581307 , 907262 , 471659 1581307 , 907262 , 471659


h 1 , h 2 ,h 3

0.33789, 0.40665, 0.50573

0.33789, 0.40665, 0.50573

r 21

1.2034

1.2034

r 32

1.2437

1.2437

f1

0.1235

0.0754

f2

0.1229

0.0757

f3
p

0.1203
6.3191

0.0802
12.2402

f ext 21

0.1237

0.0754

f ext 32

0.1237

0.0754

e a 21

0.4822%

0.4005%

32

ea

2.0841%

5.9756%

21

0.2165%

0.0463%

e ext 32

2.7672%

6.4494%

GCI fine 21

0.2712%

0.0579%

e ext

Table C-8:
Calculation of Discretization error for HPV fairing with ground clearance of 18 inches

170

30 inches
CD

CL

N 1 , N 2 , N 3 1683200, 1151200, 673160 1683200, 1151200, 673160


h 1 , h 2 ,h 3

0.33529, 0.38055, 0.45509

0.33529, 0.38055, 0.45509

r 21

1.1350

1.1350

r 32

1.1959

1.1959

f1

0.0951

0.0677

f2

0.0920

0.0683

f3
p

0.0871
0.9257

0.0735
11.2981

f ext 21

0.0955

0.0675

f ext 32

0.0928

0.0675

e a 21

3.1692%

0.8980%

e a 32

5.3280%

7.6541%

e ext 21

0.4822%

0.2832%

6.0929%

8.9288%

1.8648%

0.3529%

e ext

32

GCI fine

21

Table C-9:
Calculation of Discretization error for HPV fairing with ground clearance of 30 inches

171

Freestream
CD

CL

N1 , N2, N3

1311136, 855120 , 520484 1311136, 855120 , 520484

h 1 , h 2 ,h 3

0.59713, 0.50606, 0.43886 0.59713, 0.50606, 0.43886

r 21

1.1531

1.1531

r 32

1.1800

1.1800

f1

0.0926

0.0677

f2

0.0894

0.0683

f3
p

0.0867
0.3135

0.0735
19.2510

f ext 21

0.0927

0.0652

f ext 32

0.0895

0.0652

e a 21

3.4692%

0.3102%

e a 32

2.9480%

7.6644%

e ext 21

0.1494%

0.0214%

3.0712%

8.0214%

2.1199%

0.0267%

e ext

32

GCI fine

21

Table C-10:
Calculation of Discretization error for HPV fairing in freestream

172

Vous aimerez peut-être aussi