+ All Categories
Home > Documents > [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San...

[American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San...

Date post: 15-Dec-2016
Category:
Upload: rifki
View: 212 times
Download: 0 times
Share this document with a friend
13
American Institute of Aeronautics and Astronautics 1 Preliminary RANS Simulations and Experimental Study of a Simplified Tractor/Trailer Geometry Subrahmanya P Veluri * and Christopher J. Roy Auburn University, Auburn, Alabama, Zip 36830 Anwar Ahmed and Rifki Rifki § Auburn University, Auburn, Alabama, Zip 36830 Abstract Preliminary results are presented for Steady-state Reynolds-Averaged Navier-Stokes (RANS) simulations of the flow in the empty wind tunnel. Also presented are the experimental results of flow visualization over the simplified tractor/trailer geometry conducted at Auburn University. The simulations of the empty wind tunnel are carried out to determine the position of the simplified tractor/trailer geometry in the test section of the wind tunnel relative to the test section floor. It is observed from the computational simulations that the boundary layer thickness on the test section floor at the beginning of the test section is 0.3 inches and at the position where the trailer ends is 0.9 inches. The computational results of the empty wind tunnel are compared with the experimental data for validation. The comparisons include the boundary layer properties on the test section floor and the flow angularity calculations at the beginning of the test section. Some preliminary experiments were conducted on the simplified tractor/trailer geometry placed in the wind tunnel. A symmetric flow pattern moving outward from the center of the model is observed on the front side of the tractor. As the flow accelerated past the corner, it is observed that the flow undergoes laminar separation followed by a turbulent reattachment. Flow on the top of the tractor and also the trailer was observed to be smooth. Nomenclature C D = drag coefficient D = drag force S = cross-sectional area of the truck k = turbulent kinetic energy ω = specific dissipation rate δ = boundary layer thickness δ* = displacement thickness θ = momentum thickness ρ = density I. Introduction HE trucking industry is the backbone of the freight transportation system in the United States. According to 2003 data collected by the U.S. Department of Energy, 1 there are approximately 2.2 million tractor-trailers operating on U.S. highways. These vehicles average 62,900 miles traveled per year at a fuel consumption rate of 5.2 * Research Assistant, Aerospace Engineering Dept., 211 Aerospace Engineering Bldg., Student Member AIAA. Assistant Professor, Aerospace Engineering Dept., 211 Aerospace Engineering Bldg., Senior Member AIAA. Associate Professor, Aerospace Engineering Dept., 211 Aerospace Engineering Bldg., Senior Member AIAA. § Research Assistant, Aerospace Engineering Dept., 211 Aerospace Engineering Bldg., Student Member AIAA. T 24th Applied Aerodynamics Conference 5 - 8 June 2006, San Francisco, California AIAA 2006-3857 This material is declared a work of the U.S. Government and is not subject to copyright protection in the United States.
Transcript
Page 1: [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San Francisco, California ()] 24th AIAA Applied Aerodynamics Conference - Preliminary

American Institute of Aeronautics and Astronautics

1

Preliminary RANS Simulations and Experimental Study of a Simplified Tractor/Trailer Geometry

Subrahmanya P Veluri* and Christopher J. Roy† Auburn University, Auburn, Alabama, Zip 36830

Anwar Ahmed‡ and Rifki Rifki§ Auburn University, Auburn, Alabama, Zip 36830

Abstract

Preliminary results are presented for Steady-state Reynolds-Averaged Navier-Stokes (RANS) simulations of the flow in the empty wind tunnel. Also presented are the experimental results of flow visualization over the simplified tractor/trailer geometry conducted at Auburn University. The simulations of the empty wind tunnel are carried out to determine the position of the simplified tractor/trailer geometry in the test section of the wind tunnel relative to the test section floor. It is observed from the computational simulations that the boundary layer thickness on the test section floor at the beginning of the test section is 0.3 inches and at the position where the trailer ends is 0.9 inches. The computational results of the empty wind tunnel are compared with the experimental data for validation. The comparisons include the boundary layer properties on the test section floor and the flow angularity calculations at the beginning of the test section. Some preliminary experiments were conducted on the simplified tractor/trailer geometry placed in the wind tunnel. A symmetric flow pattern moving outward from the center of the model is observed on the front side of the tractor. As the flow accelerated past the corner, it is observed that the flow undergoes laminar separation followed by a turbulent reattachment. Flow on the top of the tractor and also the trailer was observed to be smooth.

Nomenclature CD = drag coefficient D = drag force S = cross-sectional area of the truck k = turbulent kinetic energy ω = specific dissipation rate δ = boundary layer thickness δ* = displacement thickness θ = momentum thickness ρ = density

I. Introduction HE trucking industry is the backbone of the freight transportation system in the United States. According to 2003 data collected by the U.S. Department of Energy,1 there are approximately 2.2 million tractor-trailers

operating on U.S. highways. These vehicles average 62,900 miles traveled per year at a fuel consumption rate of 5.2

*Research Assistant, Aerospace Engineering Dept., 211 Aerospace Engineering Bldg., Student Member AIAA. †Assistant Professor, Aerospace Engineering Dept., 211 Aerospace Engineering Bldg., Senior Member AIAA. ‡Associate Professor, Aerospace Engineering Dept., 211 Aerospace Engineering Bldg., Senior Member AIAA. §Research Assistant, Aerospace Engineering Dept., 211 Aerospace Engineering Bldg., Student Member AIAA.

T

24th Applied Aerodynamics Conference5 - 8 June 2006, San Francisco, California

AIAA 2006-3857

This material is declared a work of the U.S. Government and is not subject to copyright protection in the United States.

Page 2: [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San Francisco, California ()] 24th AIAA Applied Aerodynamics Conference - Preliminary

American Institute of Aeronautics and Astronautics

2

miles/gallon, resulting in an estimated consumption of 26 billion gallons of diesel fuel per year. With current diesel fuel costs near $3.00/gallon, this translates into an annual cost of $80 billion. In addition to the high costs associated with transporting goods, the U.S. produces only 40% of the oil supplied to refineries. The remaining 60% is imported from other countries, with nearly half of all imports coming from the Organization of the Petroleum Exporting Countries (OPEC).

At typical highway speeds, roughly 60% of the truck engine’s energy output goes to overcoming aerodynamic drag.2 This is due to the fact that aerodynamic drag increases as the square of the vehicle speed, while the rolling resistance between the tires and the road increase linearly with the speed. Because it is such a large portion of the engine energy output at highway speeds, reductions in aerodynamic drag can significantly reduce the vehicle’s fuel consumption. For example, a 25% reduction in the aerodynamic drag translates into a roughly 10% decrease in fuel consumed. When applied across the entire trucking industry, a 10% increase in fuel efficiency would save 2.6 billion gallons of diesel fuel per year, or approximately $8 billion. To put these numbers in perspective, if we account for the fact than only approximately half of every barrel of crude oil is used to make diesel fuel, the U.S. imported the equivalent of 37 billion gallons of diesel fuel from OPEC in 2003 (nearly half of all U.S. imports). In addition to the economic impact and the implications on oil imports, increases in fuel efficiency also translate directly into reductions in pollution emissions and are thus more environmentally friendly.

There have been a number of studies which have examined the aerodynamic drag on tractor-trailers. In the 1970s and 1980s, the majority of this work was experimental in nature. A recent review of this work was presented by Cooper3 who used both full-scale and sub-scale truck experiments to study the effects of various aerodynamic drag reduction devices for both the tractor and the trailer. More recently, researchers have also applied modern computational fluid dynamics (CFD) tools to study the aerodynamic drag of tractor-trailers. A recent DOE consortium has focused on both experimental methods and computational approaches to study the aerodynamic drag problem for trucks.2 Their study resulted in high quality experimental data at near full-scale Reynolds numbers on two different geometries: the simplified Ground Transportation System (GTS) model4 and the more realistic Generic Conventional Model (GCM).5 The experiments were designed with the dual purpose of evaluating drag reduction devices and also providing a high-quality experimental database for the validation of the CFD models. The primary modeling uncertainties are related to the choice for the turbulence model. The DOE consortium has examined both Reynolds-Averaged Navier-Stokes (RANS) approach, where all of the turbulent scales are modeled,6-8 and Large Eddy Simulation (LES), where the smaller turbulent scales are modeled but the larger scales resolved.9

Much has been learned in the last 30 years of research on aerodynamic drag reduction for tractor-trailers. Reductions in aerodynamic drag are generally reported in terms of the drag coefficient

SU

DCD2

21

∞∞

or possibly the wind-averaged drag coefficient, which accounts for fluctuations in wind velocity and direction (see Ref. 3 for details). Drag reduction techniques such as cab side-extenders and cab roof air deflectors are all commonly found on today’s tractor-trailers and have resulted in wind-averaged drag coefficient reductions of up to 0.25 from the baseline value which is near unity. More advanced techniques such as tractor-trailer gap seals and trailer side skirts are less commonly seen on U.S. highways, but also can provide significant drag reduction. The remaining region where almost no drag reduction devices are found in use is the trailer base (immediately behind the trailer). This region is not aerodynamically efficient as compared to typical aerodynamic shapes (airfoils, tear drop shapes, etc.). Storms et al.5 have shown experimentally that adding boat-tail plates or base flaps can further reduce the wind-averaged drag coefficient by 0.06; however, these add-on devices for the base region are not optimized configurations.

One way to optimize the drag reduction devices is to use CFD within some type of optimization strategy. This approach requires that the CFD tool be able to accurately predict the drag, or at least accurately predict the trends in the drag as the device is changed. The turbulence modeling approach that has the potential to produce the rapid turn-around time for drag reduction predictions is RANS, probably with wall functions used to alleviate the extremely fine wall spacing associated with integration of the turbulence modeling equations to the wall. The RANS turbulence modeling approach has been shown to accurately predict the drag for baseline configurations (i.e., without add-on base drag reduction devices); however, the details of the time averaged vortical structures and base pressure are very different from those found in experiment.6 Because the details of the time-averaged flow are not correct, it is unclear whether RANS methods will accurately predict drag or even drag trends when drag reduction devices are included. More sophisticated turbulence modeling approaches such as LES do appear to more accurately

Page 3: [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San Francisco, California ()] 24th AIAA Applied Aerodynamics Conference - Preliminary

American Institute of Aeronautics and Astronautics

3

capture the details of the flow,10 but will be much too expensive to use as the primary aerodynamic prediction tool in a drag optimization strategy.

There are a number of open questions related to aerodynamic drag on tractor-trailers. For example, it is not clear what the theoretical minimum drag coefficient is for a tractor-trailer. Standard aerodynamics packages found on U.S. trucks have a wind-averaged drag coefficient of ~0.7, while Ref. 3 indicates that additional proven technologies can further reduce this drag coefficient to ~0.55. Typical drag coefficients for airfoils can be as low as 0.01, suggesting drag coefficients for trucks still have significant room for improvement. The most sophisticated modeling approach amenable to a design optimization process requiring a large number of solutions is the steady-state RANS approach. However, the ability of RANS methods to accurately predict drag and/or drag trends has not been proven. Furthermore, it is unclear if add-on drag reduction devices can be designed on simpler shapes than full-blown tractor-trailers. Finally, even if significant advances are made in aerodynamic drag reduction, how can we ensure that the resulting designs will be cost effective and see wide-spread use by the trucking industry?

II. Program Overview Our current research efforts on tractor-trailer aerodynamics are funded by the U.S. Department of Transportation

and focus both on reducing fuel consumption (as discussed in detail above) and improving highway safety. Tractor-trailers can produce locally strong unsteady wind conditions that can be hazardous to smaller vehicles. The ultimate goal of this program is to use optimization methods to design add-on devices which reduce aerodynamic drag while at the same time reduce the large-scale fluctuation intensity in the vehicle wake. With increases in computing power, it is now becoming possible to use CFD as the aerodynamic prediction tool in a design optimization process. Part of our current research program is to demonstrate this CFD-based optimization capability.11 The other aspect of current program is to examine the validity of RANS-based turbulence models for predicting drag (or drag trends) for tractor-trailers with add-on drag reduction devices. This aspect of the program includes both wind tunnel experiments and CFD analysis of simplified tractor-trailer geometry, and is the subject of the current paper.

III. Experimental Facilities Tests were conducted in the Auburn University 3ft x 4ft test section closed circuit wind tunnel at a speed of 186

ft/sec. Prior to the tests, flow angularity in the entrance plane was measured with the help of a 5-hole probe. Presence of flow angularity due to asymmetric wind tunnel contraction as predicted by the computations was confirmed by the measurements. A flow angularity of 5 degrees along the side wall was measured. Mean flow along the tunnel centerline and the region where truck model was positioned did not exhibit any significant flow angularity. Boundary layer profiles were measured at three axial locations on the tunnel floor. At the entrance to the test section where the model nose was located, the boundary layer exhibited transitional characteristics as the flow was not tripped for forced transition. However, measurements at downstream stations showed fully turbulent profiles. Integral quantities calculated from the boundary layer profiles yielded a displacement thickness of 0.25 inches and a momentum thickness of 0.19 inches at the last measurement station. Reynolds numbers for the test based on model width were 1 million. Tunnel blockage of 8% was calculated; however, no blockage corrections were applied since these results will be used for CFD validation. Tests were conducted in the Auburn University 3ft x 4ft test section closed circuit wind tunnel capable of producing a maximum speed of 200 ft/sec. The schematics of the truck in the wind tunnel are shown in Fig. 1.

Figure 1. Drawing of the wind tunnel along with the truck geometry

Wind tunnel truck models consisted of a tractor; a short trailer and a long trailer. Figure 1 shows the tractor and

the trailer mounted in the wind tunnel. One truck model was made from solid balsawood with hardwood

Page 4: [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San Francisco, California ()] 24th AIAA Applied Aerodynamics Conference - Preliminary

American Institute of Aeronautics and Astronautics

4

reinforcements and finished with several layers of flat black paint for flow visualization purposes. The second model was equipped with pressure taps for the measurement of surface pressure distributions. Models were mounted on streamlined struts adequately positioned above the tunnel floor to prevent fully developed flow between the truck and the tunnel floor as shown in Fig. 2.

Figure 2. Simplified Tractor/Trailer model mounted in the wind tunnel test section

IV. Simplified Tractor/Trailer Geometry The simplified Tractor/Trailer geometry was based on the Modified Ground transportation System (MGTS)

geometry developed by Hammache and Browand.12 This simplified truck consists of two parts, the tractor with forward corners rounded to prevent flow separation and a rectangular trailer. For the computational simulations both the tractor and trailer are modeled together. The geometry of the simplified Tractor/Trailer which will be used for computational simulations is shown in Fig. 3. The width of the trailer is 10 inches and the height to width ratio is 1.392. Two different length trailers, with a length to width ratio of 3.4 and 4.9, are considered. The target conditions are at Reynolds numbers greater than 1 million based on the trailer width since the drag calculations and wake properties are independent of Reynolds number in this range.5

Figure 3. Isometric view of Simplified Tractor/Trailer Geometry

V. Computational Fluid Dynamics Code

A. Mesh Generation The Gridgen15 grid generation tool is used for meshing the simplified tractor/trailer geometry and the empty

wind tunnel. The wind tunnel surface data was found by taking measurements of the Auburn University wind tunnel.

Page 5: [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San Francisco, California ()] 24th AIAA Applied Aerodynamics Conference - Preliminary

American Institute of Aeronautics and Astronautics

5

The surface definition is imported in Gridgen. The test section of the wind tunnel where the truck is placed has a length of 65 inches in the axial direction and width and height of 51.25 and 36 inches, respectively. The vanes which are present in the wind tunnel for diverting the flow in the axial direction are not considered in the computational model of the wind tunnel. The vanes are replaced by a flat surface which is at a 45° angle with the axial flow direction. This surface is considered as the velocity inlet during computations, and the axial component of velocity on this boundary is defined by keeping the other two components zero. The geometry of the wind tunnel modeled with the top section cut in the horizontal plane is shown in Fig. 4.

Figure 4. Isometric view of the wind tunnel geometry

Two different meshes are created for the empty wind tunnel. The coarse mesh has approximately 1.5 million

cells and the fine mesh has approximately 4.6 million cells. Initially a structured quadrilateral mesh is used on the surface of the wind tunnel. Before generating the volume mesh, the height of the first layer of cells from the wind tunnel surface is determined such that the y+ values are close to 1 for both the meshes. A boundary layer mesh is used which has the first cell from the wind tunnel surface at a distance of 0.0005 inches for the coarse mesh and 0.00033 inches for the fine mesh with a growth rate of 1.15 and 1.097 respectively. The boundary layer region consisted of 44 and 66 layers of hexahedral cells from the wind tunnel surface for the fine and coarse meshes respectively. The remaining interior region consisted of unstructured tetrahedral cells. Empty wind tunnel simulations are run to compare computed flow angularity and boundary layer properties with the experimental data.

B. Discretization The steady-state RANS simulations are conducted on the empty wind tunnel geometry using Fluent.16 A segregated solver is used for the computations which employs a cell-centered finite volume method. A second-order upwind discretization is used for the momentum equation and a first order upwind discretization is used for turbulent kinetic energy and specific dissipation rate16. The solver settings applied in Fluent for the simulations of empty wind tunnel are tabulated in Table 1.

Table 1. Solver settings used in Fluent Solver Segregated Formulation Implicit Pressure discretization Standard Momentum discretization Second order upwind Turbulent kinetic energy First order upwind Specific dissipation rate First order upwind Pressure-Velocity coupling SIMPLE

C. Boundary conditions In the case of the empty wind tunnel, the velocities are close to Mach 0.1 and as there is not much variation in the temperatures in the wind tunnel, the flow is considered incompressible during the simulations. At the inlet, a constant stream-wise velocity boundary condition is applied. The velocity in the stream-wise direction is set to 9.1 m/s and the velocities in the other two directions are set to zero. The outlet boundary condition is set to constant pressure which is atmospheric and a gauge pressure of zero is applied. The tunnel walls are defined as stationary no-

Page 6: [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San Francisco, California ()] 24th AIAA Applied Aerodynamics Conference - Preliminary

American Institute of Aeronautics and Astronautics

6

slip walls. The boundary conditions during the simulations are applied such that the conditions match the empty wind tunnel experiments conducted at Auburn University.

The turbulence model used is the standard Wilcox 1998 k-ω, two equation model.18 The free stream turbulence parameters, k and ω are calculated using the formulae from Ref. 17. To determine these parameters, a turbulent intensity of 5% and the ratio of turbulent to laminar viscosity equal to 10 are considered. The turbulence parameters are calculated based on the average velocity in the test section of the wind tunnel.

D. Iterative convergence The convergence of the simulations is said to be achieved when all the residuals reach the required convergence

criteria. These convergence criteria are found by monitoring the percentage error in the drag. When the error in the drag becomes lower than 0.01%, then the required convergence levels are set. The convergence criterion for the continuity equation is 5E-6 and it is set to 1E-6 for the momentum, k and ω equations. The convergence of the residuals is shown in Fig. 5.

Iteration

Res

idua

ls

%E

rror

inD

rag

0 200 400 600 800 1000

10-7

10-6

10-5

10-4

10-3

10-2

10-1

100

101

10-2

10-1

100

101

102

103

ContinuityX-VelocityY-VelocityZ-Velocitykomega% Error

Figure 5. Convergence of the residuals for the empty wind tunnel simulation

VI. Results Results from the 3D computational simulations on the empty wind tunnel for the two different meshes were

obtained. Experiments were also carried out on the Auburn University wind tunnel to compare the results with the numerical predictions from the computational simulations. The empty wind tunnel results consisted of the measurements of flow angularity and the boundary layer properties. Some preliminary experiments were also conducted with the tractor/trailer model in the wind tunnel and the observations made are discussed in this section.

E. Empty Tunnel Simulations The wind tunnel consisted of a 3ft x 4ft test section which has a length of 65 inches. The upstream region of the

test section gradually changes from a rectangular cross-section to a circular cross-section as we move away from the test section. The downstream region of the test section has a gap that is opened to atmosphere. The conditions for the simulations closely matched the experimental conditions. The computational results of the empty wind tunnel are compared with the experimental results for validation. 1. Computational Predictions

A velocity of 9.1 m/s is used at the inlet to achieve an approximate average velocity of 56 m/s in the wind tunnel test section. The velocity achieved is close to the velocity measured during the experiments. The contour plots of the three components of velocity are plotted at the beginning of the test section to study the flow characteristics as the flow enters the test section. The contour plots of the three components of velocity are shown in Fig. 6, Fig. 7, and Fig. 8. It is observed that the variation of the velocity components in the inviscid core is small. The variation of the stream wise component of velocity in the inviscid region of the flow is roughly 2 m/s as shown in Fig. 6. The

Page 7: [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San Francisco, California ()] 24th AIAA Applied Aerodynamics Conference - Preliminary

American Institute of Aeronautics and Astronautics

7

variation of vertical component of velocity is slightly higher than 4 m/s and the variation of side velocity component is close to 4 m/s as shown in Fig. 7 and Fig. 8, respectively. The non-uniformities in the components of velocity are caused due to the geometry of the wind tunnel which is not symmetric in the upstream region of the flow.

Z, in

Y,i

n

-20 0 20-20

-10

0

10

20

X Velocity

5655.55554.55453.55352.55251.551

Contour plot of X-velocity at the beginning of the test section

Figure 6. Contour plot of x-velocity at the beginning of the test section

Z, in

Y,i

n

-20 0 20-20

-10

0

10

20 Y Velocity43.532.521.510.50

-0.5-1-1.5-2-2.5-3-3.5-4

Contour plot of Y-velocity at the beginning of the test section

Figure 7. Contour plot of y-velocity at the beginning of the test section

Page 8: [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San Francisco, California ()] 24th AIAA Applied Aerodynamics Conference - Preliminary

American Institute of Aeronautics and Astronautics

8

Z, in

Y,i

n

-20 0 20-20

-10

0

10

20Z Velocity

54.543.532.521.510.50

-0.5-1

Contour plot of Z-velocity at the beginning of the test section

Figure 8. Contour plot of z-velocity at the beginning of the test section

It is desirable to know the boundary layer height on the floor of the test section to determine the position of the

truck relative to the test section floor. In the case of the truck on road, there will be no boundary layer developed on the road. In the computational simulations and the experiments, a moving ground plane is not employed. Considering the boundary layer developed on the floor of the test section and the bottom side of the truck, a certain distance needs to be maintained between the truck bottom surface and the floor of the test section such that the boundary layers will not merge. The merging of the boundary layers leads to fully developed flow under the truck and can affect the wake structure behind the truck. Empty wind tunnel simulations are carried out to find the boundary layer height from the test section floor. The boundary layer height on the floor at the beginning of the test section i.e. the front end of the truck, the end of the tractor and the end of the trailer are reported. To maintain consistency with the experimental measurements, the boundary layer height is considered to be the height where the velocity reaches 95 percent of the edge velocity. At the beginning of the test section the boundary layer height is predicted to be 0.3 inches. At 20 inches from the beginning of the test section where the tractor part of the geometry ends, the boundary layer height is 0.54 inches, and it is close to 0.9 inches at a distance of 54 inches from the beginning of the test section, where the trailer ends. The velocity profiles in the boundary region at these cross sections are compared with the experimental data and are shown in Fig. 9, Fig. 10, and Fig. 11.

U/Ue

Y/δ

0 0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

1.2

CFDExperimental

Comparison of Boundary layer profiles at thebeginning of the test section, At X = 0CFD vs Experiments

Figure 9. Comparison of the velocity profiles in the boundary layer at the beginning of the test section

Page 9: [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San Francisco, California ()] 24th AIAA Applied Aerodynamics Conference - Preliminary

American Institute of Aeronautics and Astronautics

9

Figure 10. Comparison of the velocity profiles in the boundary layer at the end of the tractor

U/Ue

Y/δ

0 0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

1.2

CFDExperimental

Comparison of Boundary layer profiles at theend of the trailer, At X=54 inchesCFD vs Experiments

Figure 11. Comparison of the velocity profiles in the boundary layer at the end of the trailer

The numerical predictions of velocity profiles in the boundary layer region deviate from the experimental results at the beginning of the test section and the end of the tractor. There is a better agreement of the numerical predictions with the experimental results at the end of the trailer. These numerical results presented here are for the coarse mesh. The results of the fine mesh are close to the coarse mesh results. Their comparison with each other along with the experiments is presented in Table 2.

The flow angularity in the empty wind tunnel at the beginning of the test section predicted using the computational simulation is compared with the experimental measurements to validate the computational code. The contours plot of flow angularity at the beginning of the test section is shown in Fig. 12. The contour plot shows that the variation in the flow angularity is small at the center in major part of the flow. The variation is -1 to 2 degrees in major part of the flow. The flow angularity reaches close to 5 degrees at the top-left and bottom-left corners.

U/Ue

Y/δ

0 0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

1.2

CFDExperimental

Comparison of Boundary layer profiles at theend of the tractor, At X = 20CFD vs Experiments

Page 10: [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San Francisco, California ()] 24th AIAA Applied Aerodynamics Conference - Preliminary

American Institute of Aeronautics and Astronautics

10

Z, in

Y,i

n

-20 -10 0 10 20 30

0

10

20

30

40

Theta

54.543.532.521.510.50

-0.5-1

Contour plot of flow angularity at the beginning of the test section

Figure 12. Contour plot of flow angularity at the beginning of the test section

2. Model Validation

The numerical results are compared with the experimental results for the validation of the model. The boundary layer thickness along with the displacement thickness and momentum thickness at three different places in the wind tunnel test section floor are compared with the experimental results. The three positions involve the beginning of the tractor (beginning of the test section), end of the tractor and end of the trailer. The numerical results for both the coarse mesh and fine mesh along with the experimental results are presented Table 2.

The numerical results of the boundary layer properties, for the coarse and fine meshes closely match with each other, but predict smaller properties than those seen in the experiment. The reason for the deviation is currently under investigation.

Table 2. Comparison of numerical results of boundary layer properties with experimental results

Numerical results Boundary layer properties (inches) Coarse Fine Experimental results

δ95 0.301 0.304 0.35 δ* 0.0537 0.0544 0.0687 At the beginning of the test section

X = 0 θ 0.039 0.0394 0.0517 δ95 0.544 0.556 0.68 δ* 0.1 0.1028 0.1512 At the end of the tractor

X = 20 inches θ 0.0736 0.0753 0.112 δ95 0.9 0.915 1.22 δ* 0.1593 0.1636 0.251 At the end of the trailer

X = 54 inches θ 0.1194 0.122 0.189

The flow angularity measurements from the experiments are compared with the predicted numerical data at the beginning of the test section. The predictions along the vertical cross-section, 6 inches to the left from the center, at the center and 6 inches to the right from the center are compared with the experimental data, and are shown in Fig. 13, Fig. 14 and Fig. 15, respectively. The experimental data have error bars of ± 2.5 degrees because of the type of equipment used for measurements.

Page 11: [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San Francisco, California ()] 24th AIAA Applied Aerodynamics Conference - Preliminary

American Institute of Aeronautics and Astronautics

11

Flow Angularity, degrees

Y,i

n

-4 -2 0 2 4 6 8 10

0

5

10

15

20

25

30

35

40

ExperimentalNumerical

Flow angularity comparison at Z = - 6 inches

Figure 13. Comparison of flow angularity data at Z = -6 inches at the beginning of the test section

Flow Angularity, degrees

Y,i

n

-6 -4 -2 0 2 4 6 8

0

5

10

15

20

25

30

35

40

ExperimentalNumerical

Flow angularity comparison at Z = 0 (Vertical center)

Figure 14. Comparison of flow angularity data at the vertical center at the beginning of the test section

Page 12: [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San Francisco, California ()] 24th AIAA Applied Aerodynamics Conference - Preliminary

American Institute of Aeronautics and Astronautics

12

Flow Angularity, degrees

Y,i

n

-6 -4 -2 0 2 4 6 8

0

5

10

15

20

25

30

35

40

ExperimentalNumerical

Flow angularity comparison at Z = 6 inches

Figure 15. Comparison of flow angularity data at Z = 6 inches at the beginning of the test section

F. Flow Visualization Results of Simplified Tractor/Trailer Geometry A mixture of engine oil, yellow tempera powder, kerosene oil and oleic acid was brushed on the model and flow

was quickly brought to the test conditions. The flow pattern was established within ten minutes of run time and was photographed for later analysis.

Flow visualization results presented in Fig. 16a show the front view of the truck. Trajectory of the limiting streamlines suggested a symmetric flow pattern moving outward from the center of the model. As the flow accelerates past the corner, it undergoes laminar separation that is followed by turbulent reattachment. A band of accumulated dye near the front end of the model is indicative of a laminar separation bubble. The bubble faded and merged with the oncoming flow near the model top where the turning of streamlines was more gradual. The flow was also observed to have a slightly downward trend on the side walls of the truck as shown in Fig. 16b, which is attributed to the relaxation of the flow from the curvature encountered in the lower region.

The flow on the top of the model was observed to be smooth and uniform as shown in Fig 16c. With the addition of the trailer, trends observed remained unchanged. It may be noted that flow patterns are likely to be different when the boundary layer is force to become turbulent (to be tested in the future).

a) b) c) Figure 16. Flow visualization result in different angles a) front view, b) side view c) top view.

VII. Conclusions RANS simulations were performed on the empty wind tunnel geometry. The boundary layer height in the wind

tunnel test section is calculated to determine the position of the truck geometry relative to the test section floor. The boundary layer properties on the floor along the test section length and the flow angularity at the beginning of the

Page 13: [American Institute of Aeronautics and Astronautics 24th AIAA Applied Aerodynamics Conference - San Francisco, California ()] 24th AIAA Applied Aerodynamics Conference - Preliminary

American Institute of Aeronautics and Astronautics

13

test section are predicted and compared with the experiments carried out at Auburn University for validation. The boundary layer properties calculated from the experiments are higher than the computational predictions. After the calculation of the boundary layer height from the test section floor, the tractor/trailer geometry is placed at a height of 2 inches from the test section floor.

Some preliminary experiments were conducted on the simplified tractor/trailer geometry placed in the wind tunnel with flow visualization results being presented. A symmetric flow pattern moving outward from the center of the model is observed on the front side of the tractor. As the flow accelerates past the corner, it is observed that the flow undergoes laminar separation followed by a turbulent reattachment. Flow on the top of the tractor and also the trailer was observed to be smooth. Future plans include RANS simulations on the simplified tractor/trailer geometry positioned in the wind tunnel and the results will be compared with the experimental results.

Acknowledgements We would like to thank Andy Weldon of Auburn University for his invaluable help in building the truck model

and his aid with the experimental set up. This work is supported by Department of Transportation.

References 1. US DOE Transportation Energy Data Book: Edition 23, 2003, http://www-cta.ornl.gov/data/ 2. R. C. McCallen, K. Salari, J. M. Ortega, L. J. DeChant, B. Hassan, C. J. Roy, W. D. Pointer, F. Browand, M.

Hammache, T.-Y. Hsu, A. Leonard, M. Rubel, P. Chatalain, R. Englar, J. Ross, D. Satran, J. T. Heineck, S. Walker, D. Yaste, and B. Storms, “DOE’s Effort to Reduce Truck Aerodynamic Drag – Joint Experiments and Computations Lead to Smart Design,” AIAA Paper 2004-2249, 34th AIAA Fluid Dynamics Meeting, Portland, OR, June 2004.

3. Cooper, K. R., 2003, “Truck Aerodynamics Reborn – Lessons from the Past,” SAE Paper 2003-01-3376. 4. Storms, B. L., Ross, J. C., Heineck, J. T., Walker, S. M., Driver, D. M., and Zilliac, G. G., 2001, “An Experimental

Study of the Ground Transportation System (GTS) Model in the NASA Ames 7- by 10-ft Wind Tunnel,” NASA TM-2001-209621.

5. B. Storms, D. Satran, J. Heineck and S. Walker, “A Study of Reynolds Number Effects and Drag-Reduction Concepts on a Generic Tractor-Trailer,” AIAA-2004-2251, June 28-1, 2004.

6. C. J. Roy, J. L. Payne, and M. A. McWherter-Payne, “RANS Simulations of a Simplified Tractor/Trailer Geometry,” to appear in the ASME Journal of Fluids Engineering, September 2006.

7. Salari, K., Ortega, J. M., and Castellucci, P. J., 2004, “Computational Prediction of Aerodynamic Forces for a Simplified Integrated Tractor-Trailer Geometry,” AIAA Paper 2004-2253.

8. W. Pointer, “Evaluation of Commercial CFD Code Capabilities for Prediction of Heavy Vehicle Drag Coefficients,”AIAA-2004-2254, June 28-1, 2004

9. C. J. Roy, J. C. Brown, L. J. DeChant, and M. A. Barone, “Unsteady Turbulent Flow Simulations of the Base of a Generic Tractor/Trailer,” AIAA Paper 2004-2255, June 2004.

10. Unaune, S. V., Sovani, S. D., and Kim, S. E., “Aerodynamics of a Generic Ground Transportation System: Detached Eddy Simulation,” SAE Paper 2005-01-0548.

11. Doyle, J. B., Hartfield, R. J., and Roy, C. J., “Tractor Trailer Optimization by a Genetic Algorithm with CFD,” AIAA Paper 2006-3863, June 5-8, 2006.

12. Hammache, M., Browand, F., “On the Aerodynamics of Tractor-Trailers,” The Aerodynamics of Heavy Vehicles: Trucks, Buses and Trains, edited by R. C. McCallen, F. Browand, and J. C. Ross, Lecture Notes in Applied and Computational Mechanics, Vol. 19, Springer-Verlag, Heidelberg, 2004.

13. Roache, P. J., Verification and Validation in Computational Science and Engineering, Hermosa Publishers, New Mexico, 1998.

14. Roy, C. J., “Review of Code and Solution Verification Procedures in Computational Simulation,” Journal of Computational Physics, Vol. 205, No. 1, 2005, pp. 131-156.

15. Gridgen User Manual, Version 15, Pointwise Inc., Forth Worth, TX. 16. Fluent 6.1 User’s Guide, Fluent Inc., Lebanon, NH, 2003. 17. C. J. Roy and F. G. Blottner, “Methodology for Turbulence Model Validation: Application to Hypersonic Flows”,

Journal of Spacecraft and Rockets, Vol. 40, No.3, 2003, pp.313-325. 18. D. C. Wilcox, “Turbulence Modeling for CFD,” 2nd Edition, DCW Industries, Inc., La Canada, California, 1998.


Recommended