Reynolds Number Impact On Commercial Vehicle Aerodynamics and Performance

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 84

2015

L. RAY BUCKENDALE
LECTURE
2015-01-2859

Reynolds Number Impact on Commercial


Vehicle Aerodynamics and Performance

Richard Wood
SOLUS-Solutions and Technologies LLC

Presented at:
SAE 2015 Commercial Vehicle Engineering Congress
Rosemont, Illinois, USA
Wednesday, October 7, 2015
L. RAY BUCKENDALE LECTURE

This lecture, presented annually at the SAE Commercial Vehicle Engineering Congress, focuses on
automotive ground vehicles for either on- or off-road operation in either commercial or military service.
The intent is to provide procedures and data useful in formulating solutions in commercial vehicle design,
manufacture, operation and maintenance.

Established in 1953, this lecture commemorates the contributions of L. Ray Buckendale, 1946 SAE
President. L. Ray Buckendale, by his character and work, endeared himself to all who were associated
with him. Foremost among his many interests was the desire to develop the potential abilities in young
people. As he was an authority in the theory and practice of gearing, particularly as applied to automotive
vehicles, it was in this field that he was best able to accomplish his purpose. To perpetuate his memory,
SAE established this lecture to provide practical and useful technical information to young people involved
in vehicle engineering.

The individual presenting the Buckendale Lecture will receive an honorarium and a framed certificate. All
those attending the lecture receive an electronic copy of the lecture, and are invited to attend a sponsored
breakfast with the lecture immediately following.

2015 LECTURER

Richard Wood

Richard Wood is president and owner of SOLUS, a service-disabled veteran-owned aerodynamic


technology company. Richard has 35 years of experience in aerodynamic research and product
development for various organizations including NASA, DARPA, DOD, DOE, EPA, NASA, Boeing,
Lockheed Martin and Northrop Grumman. His extensive ground vehicle experience includes experimental
and computational fluid dynamics and fuel economy testing. Richard holds a Bachelor of Science in
Mechanical Engineering from Old Dominion University and has authored 120 technical papers with SAE,
AIAA, ASME, and NASA and has 17 patents and 7 patents pending. Mr. Wood is a registered Professional
Engineer and has received numerous awards including; 2011 Hampton Roads Entrepreneur of the Year
Award, NASA 2005 Richard Whitcomb Technology Transfer Award, NASA Engineering and Safety Center
X43A Award, and Boeing Hornet Wing Drop Program Award.

Mr. Wood is an active member of SAE International. His support for SAE meetings include; chair of the
ComVEC Commercial Vehicle Aerodynamics Committee and past chair of the ComVEC Total Vehicle
Committee where he has organized and chaired more than 30 technical sessions and reviewed more than
70 papers. In the area of SAE standards Mr. Wood is vice chair of the Truck and Bus Council, chair of the
Total Vehicle Steering Committee, and past chair of the Aerodynamics and Fuel Economy Committee. He
currently chairs the J3015 Reynolds Number task force and has chaired J1321 and J2971 task forces. In
addition to his leadership roles he is a member of the Aerodynamics and Fuel Economy Committee and
Road Vehicle Aerodynamics Forum Committee. Mr. Wood has received SAE International’s 2014 James M.
Crawford Technical Standards Board Outstanding Achievement Award and SAE International’s 2015 Forest
R. McFarland Award for his contributions to the SAE Transportation Standards Board and SAE Engineering
Meetings Board, respectively.
60TH ANNUAL L. RAY BUCKENDALE LECTURE

Reynolds Number Impact on Commercial Vehicle Aerodynamics and Performance

Richard Wood
2015-01-2859

This SAE technical paper presents the 60th Annual L. Ray Buckendale Lecture, which was
presented during the SAE Commercial Vehicle Engineering Congress on October 7, 2015 in
Rosemont, Illinois, USA.

A special “thank you” goes out to the sponsors, organizers, chairperson, and especially the
author who made this publication possible.

Corporate Sponsors
Sponsorship for the lecture is rotated among companies within the commercial vehicle
industry. The current sponsors are:

Cummins Inc.
Dana Corporation
Eaton Corporation
Meritor, Inc.
ZF TRW

Committee Members:
Steven Wesolowski, Chair Dana Corporation
Ken R. Anderson Eaton Corporation
Elizabeth Carey Cummins Inc.
Christopher Keeney Meritor, Inc.
Kevin Tilton ZF TRW

Members-at-Large:
Mehdi Ahmadian Virginia Polytechnic Institute and State University
Vern A. Caron Caron Engineering
Richard J. Hanowski Virginia Tech Transportation Institute
Donald Stanton Cummins Inc.
Daniel E. Williams ZF TRW
Mark P. Zachos DG Technologies
2015-01-2859
Published 09/29/2015
Copyright © 2015 SAE International
doi:10.4271/2015-01-2859
saecomveh.saejournals.org

Reynolds Number Impact on Commercial


Vehicle Aerodynamics and Performance
Richard Wood
SOLUS-Solutions and Technologies LLC

ABSTRACT
The impact of Reynolds number on the aerodynamics and operational performance of commercial vehicles is discussed. All supporting
data has been obtained from published experimental and computational studies for complete vehicles and vehicle components.

A review of Reynolds number effects on boundary layer state, unsteady and steady flow, time dependent wake structure, interacting
shear layer and separated flows is presented. Reynolds number modeling and simulation criteria that impact aerodynamic
characteristics and performance of a commercial vehicle are shown. The concepts of dimensional analysis and flow similarity are
employed to show that aerodynamics of commercial ground vehicles is only dependent on Reynolds number. The terminology of
Roshko is adopted for discussing the variation in drag with Reynolds number in which the subcritical, transitional and transcritical flow
regimes are defined for commercial vehicles. Criteria for aerodynamic simulation as well as testing and design of commercial vehicles
are defined and show a minimum transcritical Reynolds number value of 3 million is recommended for simulating the aerodynamics of
full-scale commercial vehicles.

Guidance is provided for aerodynamic design and analysis in the framework of Reynolds number and boundary layer flows as they
relate to previous designs, current testing and analysis criteria, and the development of future vehicles. A discussion of aerodynamic
tools is presented which call for the continued development of aerodynamic test and analysis tools capable of capturing Reynolds
number sensitivities. Of particular importance is the maturation of computational simulation tools with improved friction modeling
capability specifically developed for commercial ground vehicles.

CITATION: Wood, R., "Reynolds Number Impact on Commercial Vehicle Aerodynamics and Performance," SAE Int. J. Commer. Veh.
8(2):2015, doi:10.4271/2015-01-2859.

CHAPTER 1. INTRODUCTION disturbances from traffic and weather. Compared to the ideal test
case, these factors can dominate the aerodynamic drag of commercial
1.1. Objectives vehicles. In addition the freight industry requires aerodynamic
The commercial trucking industry is the keystone of our economy but solutions that are operationally practical, aerodynamically robust, and
continues to be burdened by the cost of fuel, the most volatile satisfy all safety requirements. It is within this context that Reynolds
operational expense. Complicating the equation are the growing number and boundary layer sensitivity factors will influence
environmental-based regulatory requirements that are driving up aerodynamic design and vehicle performance.
vehicle cost and maintenance requirements. Both the industry and
regulators have recognized aerodynamic drag reduction as a viable This 60th L. Ray Buckendale paper will build upon the 33rd lecture by
solution to these challenges. The majority of effort to date has Richard Drollinger entitled “Heavy Duty Truck Aerodynamics” [1].
focused on 1950 based aerodynamic fairing concepts where test data Drollinger's 1987 paper provides an excellent introduction to heavy
suggest reductions in aerodynamic drag up to 40 percent may be truck aerodynamics and review of the available testing and analysis
possible with corresponding improvements in fuel economy procedures of the time. His insight into the benefit of aerodynamics to
approaching 20 percent. However these projected improvements have improve the performance of commercial vehicles is reflected in the
not been consistently realized under operational conditions due to the following statements taken from his paper.
narrow design window of the aerodynamic system and the significant
difference between the operational and test environment. “In the early sixties, it was estimated that aerodynamic drag could be
reduced by 37 percent,”
To minimize the uncertainty in operational performance the
aerodynamic process must include a broad range of operational
conditions and take into account environmental factors, such as wind
590
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 591

“today it is possible to achieve over 9mpg with modern aerodynamic vehicles are performed with sub-scale models at test conditions that
shapes; high torque, low rpm engines; low profile tires and efficient may fail to satisfy the governing aerodynamic similarity criteria (i.e.
drive trains…” Reynolds number) and do not represent the critical fluid dynamic
characteristics found on a full-scale vehicle [2]. Computational
Unfortunately, after 27 years these statements remain true. An studies of representative vehicles may also fail to accurately represent
objective of this paper is to provide the industry with actionable the critical fluid dynamic characteristics found on a full-scale vehicle
aerodynamic information so that in 2040 we are not revisiting the [2]. Additionally full-scale coast down testing used to obtain
past. However, as also pointed out by Drollinger, aerodynamics is but high-speed aerodynamic data are performed over a broad range of
one of many disciplines that must work together. Reynolds numbers without an understanding of the change in the
off-body and on-body flow conditions [2]. To quantify these concerns
“Having good aerodynamics and a general fuel efficient truck can all a review was performed of published literature contained in archives
be made useless with poor driving habits.” of long standing organizations.

The goal of this paper is to provide a foundation to guide current and Recognizing that the Society of Automotive Engineers (SAE) is a
future aerodynamic design and analysis efforts. Specifically this leading source for commercial vehicle information a search of their
paper will focus on Reynolds number and boundary layer flows as ground vehicle publication archives was conducted to shed light on
they relate to previous designs, current testing and analysis criteria, the evolution of commercial vehicle aerodynamic information and
and the development of a pathway for future design efforts. To knowledge. The review begins with the year SAE added aeronautics
address these objectives requires a review of the past to establish the to its charter. From 1916 to 1960 there were less than 20 publications
role of fundamental aerodynamic principles on commercial vehicle per year that mention either “aerodynamic” or “wind resistance” or
design and performance. A brief history of Reynolds number and the “air resistance” and less than 10 percent of those mention Reynolds
boundary layer concept are presented in sub-sections 1.3 and 1.4 of number. Unfortunately, the majority of the Reynolds number content
this Chapter and a detailed discussion is contained in Chapter 2. was editorial in nature. There was a noticeable increase in
aerodynamic technical content after 1960.
To provide a clear historical record, all information contained in this
paper will be based solely on published, peer-reviewed, publically During the 30-year period from 1960 to 1990 the society published
available scientific and engineering documents. Listed at the end of an average of 70 aerodynamic documents per year. Perhaps
each Chapter are the references supporting the information presented. initially motivated by the Federal highway act of 1956 [6] and the
OPEC oil crisis [7, 8] of the 70s and 80s the publication rate
Portions of this paper have been willfully plagiarized from the increased significantly in the 1990s and is currently at 400
author's 2012 SAE publication “A Review of Reynolds Number aerodynamic publications per year. This trend with time is mirrored
Effects on the Aerodynamics of Commercial Ground Vehicles,” SAE in SAE aerodynamic publications that discuss both Reynolds
Int. J. Commer. Veh. 5(2): 2012, doi:10.4271/2012-01-2045. [2] number and boundary layer flows. However the archive search
shows that aerodynamic publications discussing Reynolds number
and those covering both Reynolds number and boundary layer
1.2. Background comprise less than 10 percent and 5 percent respectively of all SAE
“The Reynolds number is physically a measure of the ratio of inertia aerodynamic publications.
forces to viscous forces in a flow and is one of the most powerful
parameters in fluid dynamics.” John D. Anderson Jr., “Fundamentals To provide context to the SAE publications a review of the National
of Aerodynamics”, 1991 [3] Aeronautics and Space Agency (NASA) and American Institute of
Aeronautics and Astronautics (AIAA) archives was performed. This
There is a growing acceptance of aerodynamic drag reduction as a search found a significantly greater number of publications but
viable fuel economy improvement technology by all sectors of the similar trends with time, compared to SAE archives. Analysis of the
commercial vehicle industry [4]. Regulators have established several NASA and AIAA publications showed more than three times the
goals for class 8 trucks that range from a ten to fifteen percent percentage of aerodynamic publications discussing Reynolds number
increase in fuel economy associated with drag reduction to a fifty and boundary layer flows compared to SAE publications. Further
percent increase in freight efficiency by 2015. Analysis of United analysis of selected publications show that the Reynolds number and
States energy consumption show a twelve percent reduction in boundary layer content in SAE papers is less detailed and provides
aerodynamic drag will save 3.0 billion gallons of fuel [4] which less significant findings compared to that contained in AIAA and
equates to a reduction in emissions of 28.0 million tons. NASA publications.

Aerodynamic vehicle shaping and aftermarket drag reduction Another observation from the review is the influence of
technologies appear regularly on light, medium and heavy trucks and Computational Fluid Dynamics (CFD) in aerodynamic studies. Prior
trailers [5]. Supporting this growing interest is the increased use of to 1990, aerodynamic papers discussing the combination of Reynolds
aerodynamic testing and computational tools. However, anecdotal number and CFD were less than 1 percent of the total published.
evidence indicates that the sophistication of the tools and techniques Since 1990 CFD content is contained in more than 50 percent of all
being employed may not be improving at a sufficient rate to deliver published aerodynamic papers with Reynolds number content. The
the quality of data needed. Typically wind tunnel tests of commercial results highlight the growing attention paid to aerodynamics but they
592 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

raise concern with the growing dependence on CFD and the lack of continue to reference the 0.7 million minimum Reynolds number
detailed investigation of the fundamental aerodynamics of value and the use of a minimum Reynolds number value of 1.0
commercial vehicles. million by the California Air Resources Board (CARB) [67].

A technical review of a diverse set of relevant commercial vehicle For commercial vehicle aerodynamic technology to mature it is critical
literature [9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, that the use of all tools is based on the governing fundamental
25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, aerodynamic and fluid dynamic criteria. For subsonic aerodynamics the
43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55, 56, 57, 58, 59, 60, most critical criteria are Reynolds number and by association boundary
61] suggests that despite the rapid increase in the use of aerodynamic layer flows which taken together guide vehicle design principles and
technology the commercial vehicle industry has not fully embraced govern the use of testing and analysis tools. An historical review of the
Reynolds number scaling and boundary layer simulation methods as development of the Reynolds number and the boundary layer concept
a critical testing or analysis criteria. Influencing the slow adoption will be presented in the following sub-section of this Chapter.
may be the recommended minimum Reynolds number of 700,000
contained in the 1979 and 1981 versions of SAE J1252
Recommended Practice, SAE Wind Tunnel Test Procedure for Trucks
1.3. Developments and Early Use of Reynolds Number
and Buses [62]. The subject SAE document does not provide a Reynolds number and boundary layer flow effects are intertwined in
reference for the recommended Reynolds number. It is assumed that present day aerodynamic studies. This is especially true for matters
this value is based on legacy data from wind tunnel test of 1960 and related to drag and flow separation, which is synonymous with
1970 era commercial vehicles which are known to be insensitive to aerodynamic investigations of commercial vehicles. This section will
Reynolds number effects due their aerodynamically sharp-edges and provide a brief review of the early works of both Reynolds and
surfaces with small radii of curvature, see Anderson [9], Buckley [18, Prandtl, the founders of Reynolds number and boundary layer
19, 20, 21], Cooper [23, 24, 25] and Watkins [57, 58, 59, 60]. For concepts, respectively.
flow similarity the Reynolds number of the wind tunnel test shall
match of the operational vehicle. That leads to Reynolds numbers of The motivation for the section is captured in the following statements
0.7, 2.2 and 4.4 million for commercial vehicles driving at speeds of taken from Reynolds'1883 paper [68] and von Kármán's 1954 book
10, 30, and 60 mph, as we will see in Chapter 2. [69]. Reynolds' begins his 1883 paper with the following statements;

In 1976 Hucho [40] raised the initial concern that aerodynamic testing “The results of this investigation have both a practical and a
of commercial vehicles, with moderate levels of aerodynamic shaping, philosophical aspect.”
was being performed at Reynolds numbers below that required
obtaining representative data. In 1981 Gilhaus [35] suggested a “In their practical aspect they relate to the law of resistance to the
minimum width based Reynolds number value of 1.6 million. Gilhaus motion of water in pipes…”
was followed by Cooper's suggestion of 2.0 million in 1985 [26],
Drollinger supported a value of 2 million in 1987 [1], and Olson “In their philosophical aspect these results relate to the fundamental
offered 1.7 million in 1992 [46]. Note, the differences in the principles of fluid motion..”
recommended minimum Reynolds number values listed above may be
attributed to differences in the test facility, the characteristics of the In discussing Reynolds number von Kármán states:
specific vehicle in question or may be dictated by a specific component
on the vehicle. The use of the 0.7 million Reynolds number value well “The Reynolds number is now generally used in hydrodynamics,
into the 21st century is unfortunate given the guidance provided by the aerodynamics, hydraulics, and other sciences which have to do with
leading ground-vehicle aerodynamicists of the time. fluid flow. It works in some cases like black magic.”

The minimum Reynolds number debate restarted in 2000 with the “I said that the Reynolds number works like black magic, because in
launch of the Department of Energy (DOE) 21st Century Truck engineering one can sometimes use a similarity rule and other
Program [63]. In 2004 Storms [54] published wind tunnel test results general methods for the reduction of parameters without much of an
for a generic concept vehicle that showed the minimum Reynolds understanding of the phenomena.”
number is greater than 1 million and may exceed 3 million. Again the
industry was not swayed and the 0.7 million value remained the Clearly, both Reynolds and von Kármán are expressing the
standard until 2012 when SAE J1252 [62] was updated. At this time importance of understanding the foundational basis behind a finding.
the 0.7 million Reynolds number was removed from the document Extending this argument further, the foundational basis should be
and replaced with a Reynolds number sensitivity test procedure [64]. viewed as being inversely proportional to the simplicity and utility of
Follow on studies employing the Reynolds number sensitivity test by the finding. Dr. von Kármán is also raising concern that a symptom of
Wood in 2012 [2], Leuschen in 2013 [65] and McArthur in 2013 [66] a useful finding is that it can depress intellectual curiosity that will
clearly show that the minimum width-based Reynolds number is adversely impact the quality of ones future work.
greater than 2 million. Unfortunately, the J1252 Reynolds number test
procedure has yet to be generally adopted within the industry as This following review will make use of the recollections of Theodore
indicated by the significant number of current publications that von Kármán, as documented in his 1954 book entitled
“Aerodynamics” [69], the historical review of Osborne Reynolds by
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 593

Rott in 1990 [70] and by Jackson and Launder in 2007 [71] and the
historical review of Ludwig Prandtl by Anderson in 2005 [72] and by
Bodenschatz and Eckert in 2011 [73].

Reynolds seminal 1883 paper [68] reported on a series of experiments b. Sinuous (turbulent) flow.
on the flow in a tube. In Reynolds experiment the tube was connected
to a reservoir, as depicted in figure 1.1 taken from Reynolds' paper. Figure 1.2. Sketch of observed direct and sinuous flow [68].
The experiment looked at variations in pipe diameter, fluid viscosity
while the flow velocity in the tube was increased in small increments. However Reynolds did not use the present day terms of laminar and
Reynolds' interest was to understand the resistance and motion of turbulent but referred to the two states as “direct” and “sinuous”.
fluids in a pipe. He evaluated the resistance through careful These data were correlated with the resistance data to obtain a
measurements and determined the flow characteristics by introducing correlation parameter, which was not in a form of the dimensionless
dye into the flared entrance of the tube. parameter, Reynolds number, we use today. He calculated a
parameter “B” which was the reciprocal of the product of Reynolds
Reynolds developed this flow-visualization method as outlined in his number multiplied by a reference value of the kinematic viscosity. He
1877 publication [74]. The flow visualization studies showed at low continued his work in this area and published his second paper in
velocities the dye formed a thin line that was parallel to the tube 1895 [75] that was directed at calculating the critical Reynolds
centerline, see figure 1.2a image taken from Reynolds' paper. This number that he labeled “K”. In so doing he developed the Reynolds
indicated that the flow was steady and orderly. As the flow velocity averaged equations of motion.
was increased the thin line of dye thickened and became random
indicating that the flow was transitioning from orderly to random, see There is no dispute that Reynolds' 1883 [68] and 1895 [75]
lower sketch in figure 1.2b taken from Reynolds' paper. Reynolds publications are a foundational element of aerodynamics.
also noticed that as he increased the flow velocity the random motion
moved closer to the tube entrance. Regarding Reynolds' contribution, Jackson and Lauder [71] state;

“..it is no exaggeration to assert that his two principal papers on


turbulent flow essentially provided a marker for the direction of
research in Engineering Fluid Mechanics for the next century.”

However there are differing opinions as to the naming of Reynolds'


nondimensional parameter (i.e. Reynolds number). Various historians
[70, 71, 72, 73] have indicated that the naming of the parameter may
be attributed to Ludwig Prandtl, Heinrich Blasius, and Lord Rayleigh.
Rott [70] writes that Lord Rayleigh in 1892 [76] was the first to
reference Reynolds' similarity law, however von Kármán [69] states
that Sommerfield in 1908 [77] was the person who named the
parameter after Reynolds.

Von Kármán writes in 1954 [69];

“…neither Reynolds himself or other British scientists who followed


him gave a specific name to the nondimensional parameter UL/ν; it
was Arnold Sommerfield who named the parameter in honor of
Reynolds in 1908”

Regardless of who authored the “Reynolds number” label it is clear that


Reynolds' work was acknowledge and relied upon by most of those
advancing fluid mechanics, the boundary layer concept and
Figure 1.1. Sketch of Osborne Reynolds experimental apparatus [68]. aerodynamics in the 20th century. Rott points out that the adoption of
Reynolds number by the scientific community is attributed to the
writings of Prandtl in 1910 - 1914 [78, 79, 80], von Kármán in 1911
[81], and Blasius in 1912 and 1913 [82, 83]. Most notable among those
is Prandtl, the first to introduce the boundary layer concept in 1904 [84].

a. Direct (laminar) flow. In reference to Pradtl's 1904 paper, Anderson writes in 2007 [72];

“Prandtl boundary-layer idea revolutionized how scientist


conceptualized fluid dynamics.”
594 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

The linkage between Reynolds number and Prandtl's boundary layer At the time of Prandtl's 1914 findings the world was on the verge of
concept was implied in the work of Blasius in 1907 [85] and 1908 World War I and the scientific community was beginning to direct their
[86] with his investigation of boundary layer flow on a flat plate and efforts towards improving the efficiency of flight. The United States
flow separation from a circular cylinder. However, Prandtl formalized (US) lagged far behind Europe in aerodynamic understanding and in
this linkage in 1914 [79] when he correlated the change in drag with response the National Advisory Committee for Aeronautics (NACA),
Reynolds number in his investigation of Gustav Eiffel's sphere drop predecessor to NASA, was formed and SAE evolved from the Society
tests of 1912 [87]. of Automobile Engineers to the Society of Automotive Engineers, as
they included aeronautics into their charter. Both organizations focused
In 1912 Eiffel's measurements of the drag of spheres came under on power plant development during the World War I and little effort
scrutiny by Prandtl when wind tunnel data obtained in the Göttingen was directed at aerodynamics of ground vehicles. The next sub-section
wind tunnel differed with that of Eiffel by a factor of 2 [73]. To will highlight the Reynolds number and boundary-layer-based
address these concerns Eiffel performed additional testing and aerodynamic developments between 1920 and 1990 that serve as the
discovered that his original data was correct. The difference between foundation of commercial vehicle aerodynamics.
Eiffel's data and that from the Göttingen wind tunnel was the result of
a new phenomenon in which at a critical speed the drag decreased
with increasing velocity. Eiffel noted that this velocity is greater than
1.4. Review of Reynolds Number Aerodynamic
that achievable in the Göttingen wind tunnel. However, Eiffel did not Developments
offer an explanation for this effect. In 1914 Prandtl [80] investigated The fundamental findings of Reynolds and Prandtl lay the foundation
this claim by modifying the Göttingen wind tunnel and was able to for all future developments in aerodynamics. World War I clarified
reproduce Eiffel's findings. However, unlike Eiffel who correlated the the need for aeronautics research and stimulated a dramatic increase
change in drag with velocity, Prandtl correlated the drag data with in the number of wind tunnels, experimental testing, and theoretical
Reynolds number and showed that the new phenomena occurred at a developments. Ground vehicle related aerodynamic studies of the
constant Reynolds number, which can be achieved for different 1920s and 1930s followed several paths; NACA focused on
sphere sizes and different wind speeds. Prandtl offered an explanation fundamental aerodynamics, the automotive industry focused on
of this new phenomena in which he stated that the boundary layer streamlining concepts [90] while the aeronautics industry focused on
flow on the sphere is laminar at low speeds and transitions to product development. In the 1940s all aerodynamic efforts were
turbulent at a higher or critical speed. The drop in drag correlates re-directed to support World War II. During the 1950s and 1960s
with the laminar boundary layer transitioning to a turbulent boundary NACA became NASA [91] and the agency split its resources between
layer and the subsequent delay in boundary layer separation that military support and basic research while the commercial vehicle
produces a reduction in the size of the trailing wake. Representative industry published several significant studies. Driven by the 1973 oil
photographs of the phenomena are depicted in Figure 1.3 [88]. embargo [92] and 1979 energy crisis [93] the two decades from 1970
Prandtl investigated this effect further by employing a trip wire to to 1990 saw an increase in aerodynamic testing, development of new
force early transition of the boundary layer. This resulted in the drag products and the drafting of the first ever engineering test procedure
drop occurring at a lower Reynolds number. Prandtl's correlation of specifically targeted at fuel savings [94]. This sub-section will
Reynolds number, drag force, boundary layer state, flow separation, highlight significant aerodynamic and Reynolds number findings
and wake flow provides the basis for the critical elements of bluff related to commercial vehicle aerodynamics from 1920 to 2015.
body aerodynamics [89].
The 1920s and 1930s saw a significant growth in aerodynamics as
new wind tunnel facilities and the development of foundational
principles that guide current aerodynamic studies were pursued in the
US and Europe [95]. A significant effort was directed at the
development of wind tunnel test and correction methods [91].
Leading European aerodynamicists such as Hoerner, Munk, Prandtl,
and von Kármán were encouraged to continue their work in the US at
various government facilities and universities. The Daniel
Guggenheim Fund [91] provided over $3 million to eleven US
universities to develop aeronautical programs and the construction of
wind tunnels at seven of the eleven institutions. At the same time
NACA was constructing seven wind tunnels; most notable are the
Variable Density Tunnel (VDT) [94] developed by Max Munk, a
student of Prandtl, and the Full-Scale facility at the Langley
Aeronautical Laboratory [95].

The VDT allowed for testing sub-scale models at full-scale Reynolds


numbers while the Full-Scale facility allowed the testing of full size
Figure 1.3. Flow visualization of boundary layer separation and wake flow models. This unique combination of facilities generated research
about a sphere at low speed [88]. findings that defined the effect of scale and flow turbulence on test
data as well as the relationship between Reynolds number, boundary
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 595

layer flows and aerodynamic forces. Most notable are the experimental transition and separation [108, 109, 110, 111, 112]. These data
findings for simple shapes by Dryden [96, 97], Jacobs [98], Hoerner supported the development of present day computational tools in use
[99] and Lindsey [43] and the results obtained on streamwise bodies within the commercial vehicle industry.
by Munk [100, 101], Abbott [102] and Freeman [103]. In 1928 US
Navy Captain Walter Diehl captured much of these findings in his The 1950s and 1960s saw government funded aerodynamic research
book entitled “Engineering Aerodynamics”[104]. Below are several directed at transonic and supersonic aerodynamics and in 1958 space
statements from Diehl's book regarding Reynolds number. flight became the focus. There was a continuation of the fundamental
study of Reynolds number effects and boundary layer flows and the
“Reynolds Number is of such fundamental importance in influence of pressure gradients, surface curvature and roughness on
aerodynamics that a clear conception of its physical significance is transition and separation [113, 114, 115]. In this area of aerodynamics
essential. It is one of the most unusual numbers used in scientific one of the most important publication from the government labs was
work. Itself a pure number or ratio having no dimensions, it combines NASA report TN D-3579 by Braslow and Hicks [116] that defined
the effects of the four most important variables affecting air forces.” the testing standard for using grit to force boundary layer transition.
The most notable development in fundamental aerodynamics during
The following quote from Diehl is equally applicable to the this time were the publication of Schlichting's “Boundary Layer
components of a ground vehicle. Theory” in 1950 [117] and Hoerner's self-published “Aerodynamic
Drag” in 1951 [118]. Hoerner re-published his book in 1958 and 1965
“The actual value of a Reynolds Number has no significance except re-titled as “Fluid Dynamic Drag” [38].
in comparing a given series of geometrical similar forms.”…………
…..“Actually the effect of Reynolds Number is different on each item: On the application of aerodynamics to commercial vehicles the US
wings, fuselage, struts, wires, etc. Furthermore, the effect is not the and European automotive industry expanded their aerodynamic
same on all wings or on all struts, but varies widely in each group.” studies through the construction of new wind tunnels or in
partnership with universities and government labs [119, 120, 121,
Occurring in parallel with the fundamental research were wind tunnel 122] and the development of wind tunnel test methods and
studies directed at streamlining ground vehicles. The “streamlining” technology [121, 123, 124]. Each of these latter studies discusses the
design approach was not science based but may be viewed as an importance of model scale and drag sensitivity with increasing
observational based concept. Descriptions of the concept can be Reynolds number. There are several studies worth noting that address
found in several SAE papers [90, 105] in which Reynolds number commercial vehicle aerodynamics. In the 1950s the University of
and boundary layer criteria are mentioned but discounted. Despite the Maryland worked with Trailmobile to investigate tractor-trailer drag.
influence of this concept Pawlowski in 1930 [47] and of Lay in 1933 A significant finding was the drag reduction potential with edge
[106,107] investigated the effect of changes in edge radius for a rounding [125, 126], see figure 1.4.
generic bluff body. Pawlowski's and Lay's papers covered many
topics with the edge radius discussion comprising a small portion of In 1962 General Motors published their wind tunnel test findings [32]
their papers. It is unfortunate that neither paper correlated the finding on a generic tractor-trailer concept. The data also showed that edge
with Reynolds number as suggested by Prandtl in 1914 and restated rounding on the trailer front and the tractor provides significant drag
in his 1923 paper [89], entitled “The New Interpretation of the Laws reduction. These findings are consistent with those from all previous
of Resistance”. In Prandtl's 1923 paper he comments on the tests since the 1920s.
difference in drag for smooth and sharp edge shapes.

“the drag coefficient is not a constant, but a function of Reynolds


number”.

“the drag coefficient is constant over quite a long reach, especially


for bodies having edges.”

The streamlining period may have delayed fundamental aerodynamic


studies of ground vehicle. One can imagine the additional progress
that may have occurred without the distraction of streamlining.

There was minimal progress made in ground vehicle aerodynamics


during the 1940's as World War II consumed the world's attention and a. Trailer with sharp side edges
resources. Europe was once again leading the world in aerodynamic
Figure 1.4. Photographs of tractor-trailer model in the University of Maryland
knowledge and technology and in response the US moved to expand
wind tunnel [127].
its technical capability. Both Europe and the US worked to refine
wind tunnel testing capability and the US continued to expand the
understating of Reynolds number and boundary layer flows and the
influence of pressure gradients, surface curvature and roughness on
596 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

Smartway program began in 2004. This effort stimulated significant


amount of validation efforts that were redefined as verification testing.
The verification testing was based on J1252 [62] and J1321 [94].
These verification efforts grew dramatically with the launch of
CARB's mandatory aerodynamic product use in 2010 [67]. The EPA
Smartway and CARB verification efforts produced limited data on the
fluid dynamic characteristics of the subject products. Also in 2010
DOE launched the SuperTruck demonstration program with industry
[131]. At this time, limited data from these efforts have been
published. A noted exception is the 2014 publication by Smith,
Mihelic, Gifford and Ellis [132]. Perhaps the most significant
published contributions that provide an understanding of the
b. Trailer with rounded side edges fundamentals aerodynamic characteristics of commercial trucks has
been the; maturation of CFD tools, publication of SAE CFD
Figure 1.4 (cont). Photographs of tractor-trailer model in the University of guidelines J2966 [133], updates to the SAE recommended practice
Maryland wind tunnel [127].
J1252 [62] for wind tunnel testing and fuel use testing with SAE
recommended practices J1321 [134] and J [135].
It should be noted that both the General Motors test at Guggenheim
Aeronautical Laboratory at the California Institute of Technology
(GALCIT) tunnel and the University of Maryland test had difficulty 1.5. Structure of Paper
with the scale of the ground plane boundary layer affecting the force This document is structured to provide the reader with an
data. In 1967 NASA supported the automotive industry with a study understanding of the relationship between Reynolds number and
of the ground plane boundary layer in which the newly installed boundary layers and their impact on the aerodynamics and
moving-belt ground-plane in the 17 foot test section of the NASA 7 performance of commercial vehicles. Criteria for aerodynamic and
by 10 foot subsonic tunnel was used in the test of a 3/8 scale model performance simulation as well as testing and design of commercial
of a sedan. The study indicated that eliminating the floor boundary vehicles is presented. Users of this document should have a prior
layer was essential for accurate drag measurements and that a moving knowledge of aerodynamics and fluid dynamics to the level
belt was one such approach that may be used [128]. represented by Anderson [3], Hucho [39] or to have such a reference
readily available. Included in the prior knowledge is an understanding
The 1970s and 1980s brought a dramatic increase in fuel prices resulting of flow similarity. The data and information presented in this
from the 1973 oil embargo [92] and the 1979 energy crisis [93]. In document has been obtained from peer-reviewed and published
response the US established a national speed limit of 55 mph and the scientific and engineering papers and reports.
Department of Energy (DOE) was established in 1977. From 1974 to
1983 the US Department of Transportation (DOT) and SAE executed Chapter 2 reviews fundamental aerodynamic principles that impact the
the SAE/DOT Truck and Bus Fuel Economy Measurement Program aerodynamic and thereby the performance of a commercial vehicle.
[129] with contributions from the US Environmental Protection Agency Critical to this discussion are Reynolds number modeling and
(EPA), DOE and the National Research Council (NRC) of Canada. SAE simulation criteria. Included in the discussion is an overview of flow
partnered with the American Trucking Association (ATA) Technology similarity and boundary layer flows. The reader is provided the
and Maintenance Council (TMC) to develop fuel economy test necessary insight into the relationships between Reynolds number,
procedures [130]. The SAE also published a recommended practice for boundary layers, flow separation, sheer flows and wakes that is required
wind tunnel testing of trucks and busses [62]. NASA jumped into the to appreciate the complex aerodynamics of commercial vehicles.
heavy truck aerodynamic fray with fundamental shape studies on
operational vehicles [131]. The NRC initiated multiple developmental Chapter 3 introduces the aerodynamic bluff body concept, which is a
and fundamental wind tunnel investigations led by Kevin Cooper [27] class of aerodynamic shapes that includes all commercial vehicles.
and the University of Maryland continued their aerodynamic studies Bluff body vehicles have blunt fronts and truncated bases and their
[21]. General Motors sponsored a Symposium on Aerodynamic Drag aerodynamics is characterized by the presence of significant flow
Mechanisms of Bluff Bodies and Road Vehicles in 1976 [54] that separation and pressure drag which is significantly larger than friction
brought together the foremost aerodynamicist of the time. Drollinger's drag. The discussion reviews bluff body aerodynamics and presents
33rd Buckendale lecture provides a summary of the notable studies from to the reader Reynolds number and boundary layer test and analysis
1970 to 1987 [1]. criteria. The chapter presents fundamental aerodynamic trends
associated with specific geometrical features of commercial vehicles
From 1990 to 2000 fuel prices remained stable, around $1.00 a gallon, that are dependent on a combination of Reynolds number and
as did the design of most commercial vehicles. Beginning in 2000 fuel boundary layer state.
prices rose to a high of nearly $5.00 a gallon in 2008 and have
averaged approximately $3.00 a gallon since. The DOE launched the Chapters 4 and 5 provide the reader with an overview of commercial
21st Century Truck Partnership in 1998 [63]. The program produced a vehicle Computational Simulations and Experimental Testing
limited number of fundamental aerodynamic publications addressing capabilities, respectively. Each Chapter presents the critical
Reynolds number effects [10, 27, 52, 55]. The voluntary EPA simulation and testing factors that are influenced by Reynolds number
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 597

and boundary layer modeling. Information for proper modeling of the 14. Barlow, J., Rae, W. and Pope, A., “Low-Speed Wind Tunnel Testing,”
3rd Edition. John Wiley and Sons Inc., ISBN: 978-0-471-55774-6, 1999.
governing physics for all vehicle components are outlined and
15. Barnard, R., “Road Vehicle Aerodynamic Design: An Introduction,” 2nd
discussed. The reader should leave these chapters with insight into Edition. MechAero Publishing, ISBN: 0954073401, 2001.
Reynolds number effects that impact the design and evaluation of 16. Final Report, “Part I- Experimental Measurement of the Flow Field of
commercial vehicles. Furthermore the reader will be able to Heavy Trucks,” University of Southern California, sponsored by US
DOE with contract DE-AC 26-98EE50512, 2005.
recognize the limitations and applicability of a facility or method that
17. Browand, F., Radovich, C., and Boivin, M., “Fuel Savings by Means
is critical to establishing credibility of a vehicle evaluation. of Flaps Attached to the Base of a Trailer: Field Test Results,” SAE
Technical Paper 2005-01-1016, 2005, doi:10.4271/2005-01-1016.
Chapter 6 discusses the application of Reynolds number and 18. Buckley, F., Walston, W., and Marks, C., “Fuel Savings from Truck
Aerodynamic Drag Reducers and Correlation with Wind-Tunnel Data,”
boundary layer criteria to the design of current and future commercial J. Energy, Vol 2, No. 6, 1978.
vehicles such as buses, single unit trucks and combination vehicles. 19. Buckley, F., Marks, C., and Walston, W., “Analysis of Coast-Down Data
The Chapter provides a brief review of vehicle design trends over the to Assess Aerodynamic Drag Reduction on Full-Scale Tractor-Trailer
Trucks in Windy Environments,” SAE Technical Paper 760850, 1976,
last four decades highlighting the increased use of aerodynamics and doi:10.4271/760850.
aerodynamic shaping to increase vehicle performance. This trend has 20. Buckley, F. and Marks, C., “Feasibility of Active Boundary-Layer-
dramatically increased the importance of Reynolds number, boundary Control Methods for Reducing Aerodynamic on Tractor Trailer Trucks,”
Journal of Wind Eng. and Ind. Aerodynamics, 4 (2) 133-148, 1979.
layer management and flow separation control in the design process. 21. Buckley, F. and Sekscienski, W., “Comparisons of Effectiveness of
The reader should take away from this Chapter that the importance of Commercially Available Devices for the Reduction of Aerodynamic
the Reynolds number and its impact on the design process, Drag on Tractor-Trailers,” SAE Technical Paper 750704, 1975,
doi:10.4271/750704.
aerodynamic testing, viscous-pressure drag ratio, and vehicle 22. Cogotti, A., “Aerodynamic Characteristics of Car Wheels,” International
performance will continue to increase. Journal of Vehicle Design, SP 3:173-196, 1983.
23. Cooper, K., “The Wind Tunnel Testing of Heavy Trucks to
Reduce Fuel Consumption,” SAE Technical Paper 821285, 1982,
Chapter 7 provides the reader a summary of the document including a doi:10.4271/821285.
detailed list of the critical analysis, testing and design criteria 24. Cooper, K., Gerhardt, H., Whitbread, R., Garry, K., and Carr, G., “A
discussed in each Chapter. The reader will also be presented a Comparison of Aerodynamic Drag Measurements on Model Trucks in
Closed-Jet and Open-Jet Wind Tunnels,” Journal of Wind Eng. and Ind.
discussion of future directions and challenges in vehicle design, Aerodynamics, 22:299-316, 1986.
aerodynamics and Reynolds number effects.1.6. 25. Cooper, K., Mason, W., and Bettes, W., “Correlation Experience with the
SAE Wind Tunnel Test Procedure for Trucks and Buses,” SAE Technical
Paper 820375, 1982, doi:10.4271/820375.
1.6. References 26. Cooper, K., “The Effect of Front-Edge Rounding and Rear-Edge
1. Drollinger, R., “Heavy Duty Truck Aerodynamics,” SAE Technical Shaping on the Aerodynamic Drag of Bluff Vehicles in Ground
Paper 870001, 1987, doi:10.4271/870001. Proximity,” SAE Technical Paper 850288, 1985, doi:10.4271/850288.
2. Wood, R., “A Review of Reynolds Number Effects on the Aerodynamics 27. Cooper, K. and Leuschen, J., “Model and Full-Scale Wind Tunnel Tests
of Commercial Ground Vehicles,” SAE Int. J. Commer. Veh. 5(2):628- of Second-Generation Aerodynamic Fuel Saving Devices for Tractor-
639, 2012, doi:10.4271/2012-01-2045. Trailers,” SAE Technical Paper 2005-01-3512, 2005, doi:10.4271/2005-
01-3512.
3. Anderson, J., “Fundamentals of Aerodynamics,” 2nd Ed., McGraw-Hill,
ISBN: 7802434475, 1991 28. Croll, R., Gutierrez, W., Hassan, B., Suazo, J. et al., “Experimental
Investigation of the Ground Transportation Systems (GTS) Project for
4. McCallen, R., “DOE's Effort to reduce Truck Aerodynamic Drag-Joint Heavy Vehicle Drag Reduction,” SAE Technical Paper 960907, 1996,
Experiments and Computations lead to Smart Designs,” AIAA 2004- doi:10.4271/960907.
2249, 34th AIAA Fluid Dynamics Conference and Exhibit, Portland OR,
2004. 29. Damiani, F., Iaccarino, G., Kalitzin, G. and Khalighi, B., “Unsteady
Flow Simulations of Wheel-Wheelhouse Configurations,” AIAA 2004-
5. Belzile, M., “Review of Aerodynamic Drag Reduction Devices for 2344, 2004.
Heavy Trucks and Buses,” NRC-CNRC Technical Report CSTT-HVC-
TR-205. 2012. 30. Delaney, N., and Sorensen, N., “Low-Speed Drag of Cylinders of
Various Shapes,” NACA TN 3038, 1953.
6. Federal Highway Act of 1956, Public Law 627, Chapter 462, June 29,
1956. 31. Dryden, H., Schubauer, G., Mock, W. and Skramstad, H.,
“Measurements of Intensity and Scale of Wind Tunnel Turbulence and
7. Foreign Relations of the United States, 1969-1976, Volume XXXVI, Their Relation to the Critical Reynolds Number of Spheres,” NACA
Energy Crisis, 1969-1974. Editor: Linda Qaimmaqami, General Rpt. No. 581, 1937.
Editor: Edward C. Keefer, United States Government Printing Office
Washington 2011 32. Flynn, H. and Kyropoulos, P., “Truck Aerodynamics,” SAE Technical
Paper 620531, 1962, doi:10.4271/620531.
8. Foreign Relations of the United States, 1969-1976, Volume XXXVII,
Energy Crisis, 1974-1980. Editor: Steven G. Galpern, General 33. Fackrell, J. and Harvey, J., “The Aerodynamics of an Isolated Road
Editor: Edward C. Keefer, United States Government Printing Office Wheel,” In Proc. of AIAA 2nd Symposium on Aerodynamics of Sports
Washington, 2012 and Competition Automobiles, 16: 119-125, Los Angeles, CA, 1974.
9. Anderson, J., Firey, J., Ford, P., and Kieling, W., “Truck Drag 34. Gad-el-Hak, M. and Bushnell, D., “Status and Outlook of Flow
Components by Road Test Measurement,” SAE Technical Paper 640794, Separation Control,” AIAA-91-0037, 29th Aerospace Sciences Meeting,
1964, doi:10.4271/640794. Reno, NV, 1991.
10. Arcas, D., Browand, F., and Hammache, M., “Flow Structure in the Gap 35. Gilhaus, A., “The Influence of Cab Shape on Air Drag of Trucks,”
Between Two Bluff Bodies,” AIAA 2004-2250, 2004 Journal of Wind Eng. and Ind. Aerodynamics, 9(1-2):77-87, 1981.
11. Axon, L., Garry, K., and Howell, J., “The Influence of Ground Condition 36. Grunwald, K.J., “Aerodynamic Characteristics of Vehicle Bodies
on the Flow Around a Wheel Located Within a Wheelhouse Cavity,” at Crosswind Conditions in Ground Proximity,” NASA Tech. Note
SAE Technical Paper 1999-01-0806, 1999, doi:10.4271/1999-01-0806. D-5935, 1970.
12. Baker, C. and Humphreys, N., “Assessment of the Adequacy of Various 37. Croll, R., Gutierrez, W., Hassan, B., Suazo, J. et al., “Experimental
Wind Tunnel Techniques to Obtain Aerodynamic Data for Ground Investigation of the Ground Transportation Systems (GTS) Project for
Vehicles in Cross Winds,” Journal of Wind Engr. and Ind. Aerodynamics Heavy Vehicle Drag Reduction,” SAE Technical Paper 960907, 1996,
60, 49-68, 1996. doi:10.4271/960907.
13. Barlow, J., Guterres, R., and Ranzenbach, R., “Rectangular Bodies with
Radiused Edges in Ground Effect,” AIAA-99-3153, 1999
598 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

38. Hoerner, S., “Fluid-Dynamic Drag: Practical Information on 62. SAE International Surface Vehicle Recommended Practice, “Wind
Aerodynamic Drag and Hydrodynamic Resistance,” Bricktown New Tunnel Test Procedure for Trucks and Buses,” SAE Standard J1252,
Jersey, 1965 Issued August 1979.
39. Hoffman, J., Martindale, B., Arnette, S., Williams, J. et al., “Effect of 63. “Review of the 21st Centruy Truck Partnership,” National Academies
Test Section Configuration on Aerodynamic Drag Measurements,” SAE Press ISBN: 978-0-309-12208-5. 2008.
Technical Paper 2001-01-0631, 2001, doi:10.4271/2001-01-0631. 64. SAE International Surface Vehicle Recommended Practice, “Wind
40. Hucho, W.-H., “Aerodynamics of Road Vehicles, 4th Edition,” Tunnel Test Procedure for Trucks and Buses,” SAE Standard J1252,
(Warrendale, SAE International, 1998), ISBN 978-0-7680-0029-0. Rev. August 2012.
41. Hucho, W., Janssen, L., and Emmelmann, H., “The Optimization of 65. Leuschen, J., “Considerations for the Wind Tunnel Simulation of
Body Details-A Method for Reducing the Areodynamic Drag of Road Tractor-Trailer Combinations: Correlation of Full- and Half-Scale
Vehicles,” SAE Technical Paper 760185, 1976, doi:10.4271/760185. Measurements,” SAE Int. J. Commer. Veh. 6(2):529-538, 2013,
42. Kim, M. and Geropp, D., “Experimental Investigation of the Ground doi:10.4271/2013-01-2456.
Effect on the Flow Around Some Two-Dimensional Bluff Bodies with 66. McArthur, D., Burton, D., Thompson, M., and Sheridan, J.,
Moving Belt Technique,” Journal of Wind Eng. and Ind. Aerodynamics, “Development of a Wind Tunnel Test Section for Evaluation of Heavy
74-76:511-519, 1998. Vehicle Aerodynamic Drag at a scale of 1:3,” SAE Int. J. Commer. Veh.
43. Lindsey, W., “Drag of Cylinders of Simple Shapes,” NACA Report No. 6(2):522-528, 2013, doi:10.4271/2013-01-2455.
619. 1937. 67. “Implementation Guidance for the Tractor-Trailer GHG Regulation,”
44. Metz, L. and Sensenbrenner, K., “The Influence of Roughness Elements California Environmental Protection Agency, Air Resources Board,
on Laminar to Turbulent Boundary Layer Transition as Applied to Scale October 2012.
Model Testing of Automobiles,” SAE Technical Paper 730233, 1973, 68. Reynolds O., “An experimental investigation of the circumstances
doi:10.4271/730233. which determine whether the motion of water in parallel channels shall
45. Muirhead, V. and Saltzman, E., “Reduction of Aerodynamic Drag be direct or sinuous and of the law of resistance in parallel channels,”
and Fuel Consumption for Tractor Trailer Vehicles”, AIAA journal of Philos. Trans. R. Soc. 174:935-82, 1883
Energy, 2(5): 279-284, 1979. 69. von Kármán, T., “Aerodynamics: Selected Topics in the Light of
46. Olson, M. and Schaub, U., “Aerodynamics of Trucks in Wind Tunnels: their Historical Development,” Cornell University Press and Oxford
The Importance of Replicating Model Form, Model Detail, Cooling University press, 1954
System and Test Conditions,” SAE Technical Paper 920345, 1992, 70. Rott, N., “Note on the History of the Reynolds Number,” Annu. Rev.
doi:10.4271/920345. Fluid Mech, 22: I-II, 1990.
47. Pawlowski, F., “Wind Resistance of Automobiles,” SAE Journal 27(1): 71. Jackson, D. and Launder, B., “Osborne Reynolds and the Publication of
5-15, 1930. His Papers on Turbulent Flow,” Annu. Rev. Fluid Mech. 2007. 39:19-35
48. Peterson, R., “Drag Reduction Obtained by the Addition of a Boattail to 72. Anderson, J. D. Jr., “Ludwig Prandtl's Boundary Layer,” Annu. Rev.
a Box Shaped Vehicle,” NASA CR 163113, 1981. Fluid Mech. 2007, 39:19-35
49. Roshko, A. “Experiments on the flow past a circular cylinder at very 73. Bodenschatz, E. and Eckert, M., “Prandtl and The Göttingen School,”
high Reynolds number.” J. Fluid Mech. 10(3):345-356, 1961. chapter 2 of: Davidson, P. A., Kaneda, Y., Moffatt, K. and Sreenivasan,
50. Saunders, J., Watkins, S., Hoffmann, P., and Buckley, F., “Comparison K.R., “A Voyage Through Turbulence,”. Cambridge University Press,
of On-Road and Wind-Tunnel Tests for Tractor-Trailer Aerodynamic ISBN: 9780521198684, 2011.
Devices, and Fuel Savings Predictions,” SAE Technical Paper 850286, 74. Reynolds, O., “On Various Forms of Vortex Motion,” Proceedings of the
1985, doi:10.4271/850286. Literary and Philosophical Society of Manchester, Feb. 1877, reprinted
51. Schewe, G., “Reynolds Number Effects in Flow Around More-or-Less as paper 24:183-191 of:Reynolds, O., “Papers on Mechanical and
Bluff Bodies,” Journal of Wind Eng. and Ind. Aerodynamics 89(14-15): Physical Subjects”, Tome I, Cambridge University Press, 1900.
1267-1289, 2001. 75. Reynolds O., “On The Dynamical Theory of Incompressible Viscous
52. Schoon, R. and Pan, F., “Practical Devices for Heavy Truck Fluids and the Determination of the Criterion.” Philos. Trans. R. Soc.
Aerodynamic Drag Reduction,” SAE Technical Paper 2007-01-1781, 186:123-164, 1895
2007, doi:10.4271/2007-01-1781. 76. Strutt, J. W., 3rd Baron Rayleigh, “On the question of the stability of
53. Sheridan, A. and Grier, S., “Drag Reduction Obtained by Modifying a the flow of fluids,” Philos. Mag. Series 5, No. 34(206): 59-70. 1892.
Standard Truck,” NASA TM 72846, 1978. Reprinted in Scientific Papers 3:575-84, Cambridge Univ. Press, 1920.
54. Sovran, G., Morel, T. and Mason, W., eds. “Aerodynamic Drag 77. Sommerfeld, A. “Ein Beitrag zur hydrodynamischen Erklärung der
Mechanisms of Bluff Bodies and Road Vehicles,” NewYork: Plenum, turbulenten Flüssigkeitsbewegung,” 4th Int. Congr. Math., Rome, 3:116-
1978 24, 1908.
55. Storms, B., Satran, D., Heineck, J. and Walker, S., “A Study of Reynolds 78. Prandtl, L., “Eine Beziehung zwischen Wärmeaustausch und
Number Effects and Drag Reduction Concepts on a Generic Tractor Strömungswiderstand der Flüssigkeiten,” Phys. Zeitschr. 11:1072-78,
Trailer,” AIAA 2004-2251, 2004. 1910. Reprinted on pp.585-596 in:Tollmien, W., Schlichting, H. and
Görtler, H., “Ludwieg Prandtl Gesammelte Abhandlungen,” Springer
56. Veldhuis, L. and Henneman, B., “Experimental and Numerical Study of
Berlin Heidelberg, 1961.
Leading Edge Separation on Blunt Bodies,” SAE Technical Paper 2007-
01-4291, 2007, doi:10.4271/2007-01-4291. 79. Prandtl, L., “Flüssigkeitsbewegung” Chapter 3 of: Korschelt, E., Linck,
G. and Oltmanns, F., “Handwörterbuch der Naturwissenschaften,” Vol.
57. Watkins, S, Saunders, J. and Gibson, K., “Tri-Axle Tipper Trucks -
4(Fluorgruppe - Gewebe): 101-40, Gustav Fischer Verlag, 1913.
Some New and Simple Devices as Aerodynamic Means of Saving Fuel
- Road and Tunnel Tests,” 9th Australian Fluid Dynamics Conference, 80. Prandtl, L., “Der Luftwiderstand von Kugeln,” Nachrichten d. Ges. d.
Auckland, 1986. Wiss. zu Göttingen, Mathematisch-physikalische Klasse, pp. 177-190,
1914.
58. Watkins, S. and Cooper, K., “The Unsteady Wind Environment
of Road Vehicles, Part Two: Effects on Vehicle Development and 81. von Kármán, T., “Über die Turbulenzreibung verschiedener
Simulation of Turbulence,” SAE Technical Paper 2007-01-1237, 2007, Flüssigkeiten,” Phys. Z. 12:283-84, 1911. Reprinted in:“Collected Works
doi:10.4271/2007-01-1237. of Theodore von Kármán,” Edition 1, Vol. 1:321-323, Butterworth
Scientific, 1956.
59. Watkins, S., Hoffman, P. and Saunders, J., “Comparison of On-Road
and Wind-Tunnel Tests for Rigid Truck Aerodynamic Devices,” 9th 82. Blasius, H., “Das Ähnlichkeitsgesetz bei Reibungsvorgängen,” VDI-Z.
Australasian Fluid Mechanics Conference, Auckland, December 8-12, 16: 639-643, 1912.
1986. 83. Blasius, H., “Das Ähnlichkeitsgesetz bei Reibungsvorgängen in
60. Watkins, S., Saunders, J. and Hoffman, P., “Comparison of Road and Flüssigkeiten,” VDI Mitt. Forschungsarb. 131:1-41, Springer-Verlag
Wind-Tunnel Drag Reductions for Commercial Vehicles, Journal of Berlin Heidelberg, 1913.
Wind Engr. and Ind. Aerodynamics, 49 (1993) 411-420. 84. Prandtl, L. “Über Flüssigkeitsbewegung bei sehr kleiner Reibung,”
61. Zanacic, A. and Long, K., “Drag and Wake Characteristics of Three Verhandlungen des III. Internationalen Mathematiker-Kongresses,
Dimensional Bluff Bodies with Varying Radii of Curvature,” AIAA Heidelberg, pages 484-491, 1905. Reprinted on pp. 575-584 in:Tollmien,
2009-3612, 27th AIAA Applied Aerodynamics Conference, San Antonio W., Schlichting, H. and Görtler, H., “Ludwieg Prandtl Gesammelte
TX, 2009 Abhandlungen,” Vol. 2, Springer Berlin Heidelberg, 1961.
85. Blasius, H. “Grenzschichten in Flüssigkeiten mit kleiner Reibung,” PhD
Dissertation, University of Göttingen, 1907.
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 599

86. Blasius, H., “Grenzschichten in Flüssigkeiten mit kleiner Reibung,” Z. 113. Klebanoff, P and Diehl, Z., “Some Features of Artificially Thickened
Math Phys 56:1-37; 60: 397-398, 1908. Translated to english as: NACA Fully Developed Turbulent Boundary layers with Zero Pressure
TM 1256, 1950. Gradient,” NACA TR 1110, 1952. Supersedes NACA TN 2475 with the
87. Eiffel, A., “Sur la résistance des sphères dans l'air en movement,” same title and authors from 1951.
Comptes Rendues, 155:1597-1599, 1912. 114. Sandborn, V., “Preliminary Experimental Investigation of Low-Speed
88. Gelzer, C., “Fairing Well: Aerodynamic Truck research at NASA's Turbulent Boundary Layers in Adverse Pressure Gradients,” NACA TN
Dryden Flight Research Center. NASA SP-2010-4545, 2011 3031, 1953.
89. Prandtl, L. “The New Interpretation of the Laws of Resistance,” NACA 115. Braslow, A., Harris, R., and Hicks, R., “Use of Grit-Type Boundary-
TM-198, 1923 Layer Transition Trips on Wind-Tunnel Models”, NASA TN D-3579,
1966
90. Brown, L and Chase, H. “Streamlining--Up-to-Date Facts and
Developments,” SAE 340013, 1934 116. Schlichting, H., “Boundary Layer Theory”, McGraw-Hill, Inc., 1950.
Most recent: 8th revised and enlarged edition, ISBN: 3-540-66270-7,
91. Roland, A., “Model Research,” NACA SP-4103 Vol.1 : 1915-1958,
2003.
1985.
117. Hoerner, S., “Aerodynamic Drag,” self-published, 1951. Revised
92. Merrill, K., “The Oil Crisis of 1973-1974: A Brief History with
edition published as “Fluid-Dynamic Drag,” Hoerner Fluid Dynamics,
Documents,” Bedford St. Martin's, 2007
Bricktown New Jersey, 1965.
93. Verleger, P., “The U. S. Petroleum Crisis of 1979,” Brookings Papers on
118. Ludvigsen, K., “The Time Tunnel - An Historical Survey of
Economic Activity, 1979(2) : 463-476, 1979
Automotive Aerodynamics,” SAE Technical Paper 700035, 1970,
94. SAE International Surface Vehicle Recommended Practice, “SAE Joint doi:10.4271/700035.
TMC/SAE Fuel Consumption Test Procedure-Type II,” SAE Standard
J1321, Oct. 1986 119. White, J., “New Techniques for Full Scale Testing,” SAE Technical
Paper 600479, 1960, doi:10.4271/600479.
95. Baals, D. and Corliss, W., “Wind Tunnels of NASA,” NASA SP 440,
120. Kelly, K., Kyropoulos, P., and Tanner, W., “Automobile Aerodynamics,”
1981.
SAE Technical Paper 600195, 1960, doi:10.4271/600195.
96. Dryden, H and Kuethe, A., “Effect of Turbulence in Wind Tunnel
121. Esper, A., “Early Wind Tunnel Testing at Ford,” SAE Technical Paper
measurements,” NACA TR 342, 1932
600480, 1960, doi:10.4271/600480.
97. Schubauer, G. and Dryden, H., “The Effect of Turbulence on the Drag of
122. Kessler, J. and Wallis, S., “Aerodynamic Test Techniques,” SAE
Flat Plates,” NACA TR 546, 1932.
Technical Paper 660464, 1966, doi:10.4271/660464.
98. Jacobs, E., “Sphere Drag Tests in the Variable Density Wind Tunnel,”
123. Gross, D. and Sekscienski, W., “Some Problems Concerning Wind
NACA TN 312, 1929
Tunnel Testing of Automotive Vehicles,” SAE Technical Paper 660385,
99. Hoerner, S., “Tests of Spheres with Reference to Reynolds Number, 1966, doi:10.4271/660385.
Turbulence, and Surface Roughness,” NACA TM 777, 1935. Translation
of:“Versuche mit Kugeln betreffend Kennzahl, Turbulenz und 124. Brouer, O., “Future Trends in Automotive Transportation and
Oberflächenbeschaffenheit,” Luftfahrtforschung 12(1):42-54, 1935. Maintenance,” SAE J0255, pp. 74, 1955.
125. Sherwood, W., “Wind Tunnel test of Trailmobile Trailers,” University of
100. Munk, M., “The Drag of Zeppelin Airships,” NACA TR 117, 1923.
Maryland Wind Tunnel Report No. 85. College Park, MD, April 1974.
101. Munk, M., “The Aerodynamic Forces on Airship Hulls,” NACA TR 184,
126. Cooper, K., “Commercial Vehicle Aerodynamic Drag Reduction:
1924.
Historical Perspective as a Guide,” The Aerodynamics of Heavy
102. Abbott, I., “Airship Model Tests in the Variable Density Wind Tunnel,” Vehicles: Trucks, Buses, and Trains, Lecture Notes in Applied and
NACA TR 394, 1931. Computational Mechanics Volume 19:9-28, 2004.
103. Freeman, H., “Measurements of Flow in the Boundary Layer of a 127. Turner, D., “Wind-Tunnel Investigation of a 3/8-Scale Automobile
1/40-Scale Model of the U.S. Airship “Akron”,” NACA TR 430, 1932. Model Over a Moving-Belt Ground Plane,” NASA TN D-4229, 1967.
104. Diehl, W., “Engineering Aerodynamics,” David W. Taylor Naval Ship 128. Ringham, R., “Executive Summary SAE/DOT Truck and Bus Fuel
Research and Development Center, The Ronald Press Company, 1928, Economy Measurement Program,” SAE Technical Paper 831783, 1983,
Revised Edition available from 1936. doi:10.4271/831783.
105. Brown, L., “Economics of Streamlining in Heavy Transportation,” SAE 130. Johnson, R., “Development of Fuel Economy Test Procedures,” SAE
Technical Paper 360022, 1936, doi:10.4271/360022. Technical Paper 831784, 1983, doi:10.4271/831784.
106. Lay, W., “Is 50 Miles Per Gallon Possible With Correct Streamlining?,” 131. Recovery Act - Systems Level Technology Development, Integration,
SAE Technical Paper 330039, 1933, doi:10.4271/330039. and Demonstration for Efficient Class 8 Trucks (SuperTruck) and
107. Lay, W., “Is 50 Miles Per Gallon Possible With Correct Streamlining?,” Advanced Technology Powertrains For Light-Duty Vehicles (ATP-LD)
SAE Technical Paper 330041, 1933, doi:10.4271/330041. Funding Opportunity: DE-FOA-0000079 - Cooperative Agreement
108. von Doenhoff, A. and Tetervin, N, “Determination of General Relations number DE-EE0003403.
for the Behavior of Turbulent Boundary Layers,” NACA TR 772, 1943. 132. Smith, J., Mihelic, R., Gifford, B., and Ellis, M., “Aerodynamic Impact
109. Wieghardt, K. and Tillman, W., “On The Turbulent Friction Layer for of Tractor-Trailer in Drafting Configuration,” SAE Int. J. Commer. Veh.
Rising Pressure,” NACA TM 1314, 1944. 7(2):619-625, 2014, doi:10.4271/2014-01-2436.
110. Liepmann, H., “Investigation of Boundary Layer Transition on Concave 133. SAE International Surface Vehicle Recommended Practice, “Guidelines
Walls,” NACA ACR 4J28, 1945. for Aerodynamic Assessment of Medium and Heavy Commercial
111. Bussmann, K. and Ulrich, A., “Systematic Investigations of the Ground Vehicles Using Computational Fluid Dynamics,” SAE Standard
Influence of the Shape of the Profile Upon the Position of the J2966, Issued Sept. 2013.
Transition Point,” NACA TM 1185. 1947. Translation of“Systematische 134. SAE International Surface Vehicle Recommended Practice, “Fuel
Untersuchungen über den Einfluss der Profilform auf die Lage des Consumption Test Procedure - Type II,” SAE Standard J1321, Rev. 2012
Umschlagpunktes,” Technische Berichte 10(9), 1943. 135. SAE International Surface Vehicle Recommended Practice, “SAE Fuel
112. Schlichting, H., “Lecture Series “Boundary Layer Theory” Part II - Consumption Test Procedure (Engineering Method),” SAE Standard
Turbulent Flows,” NACA TM 1218, 1949. J1526, Proposed Draft July 2015.
600 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

CHAPTER 2. AERODYNAMIC Commercial ground vehicle or test article geometry is typically


FUNDAMENTALS non-dimensionalized by the vehicle width. The fluid velocity is
non-dimensionalized by the ratio of the local fluid velocity to the free
The aerodynamics of commercial vehicles are extremely complex
stream fluid velocity. Fluid density is non-dimensionalized in terms of
consisting of varying boundary layer states, unsteady and steady
the ratio of local fluid density to the free stream fluid density. Local
flows, time dependent wake structures, vortices, bubbles, interacting
pressure is non-dimensionalized in a pressure coefficient defined as
shear layers and separated flows with and without reattachment.
the difference between the local pressure (due to the presence of the
Adding to the complexity is the presence of vehicle-based thermal
body) and the free stream pressure divided by the free stream dynamic
and vibrational effects as well as atmospheric and operational based
pressure. Local temperature will be non-dimensionalized by its ratio to
turbulence. A third level of complexity is the presence of internal
the free stream value. The time will be non-dimensionalized as a ratio
flows and rotational systems. All of these fluid dynamic and
to the time for a fluid particle to travel the reference length at the
geometric characteristics are Reynolds number sensitive.
speed of flow far from the body. When time dependence is averaged
out or the actual circumstance of having steady boundary conditions
To provide context, this Chapter will review fundamental yields steady results, the aerodynamic coefficients are functions only
aerodynamic and fluid dynamic concepts that impact the aerodynamic of the dimensionless similarity parameters, Reynolds number and
characteristics and performance of a commercial vehicle. Critical to Mach number along with the orientation angle of the test article. Each
this discussion are Reynolds number modeling and simulation solution of the non-dimensional system for a value of the Reynolds
criteria. Included in the discussion is an overview of flow similarity number provides a result that applies equally to different fluids as well
and boundary layer flows. It is beneficial if the reader has prior as different sizes of bodies at differing speeds and differing
knowledge of aerodynamics and fluid dynamics to the level coefficients of viscosity as long as the similarity parameters are
represented by Anderson [3], Hucho [40] or Shames [136]. matched. For subsonic aerodynamic studies of commercial ground
vehicles the density may be assumed to be constant, as long as the
2.1. Dimensional Analyses and Dynamic (Flow) Similarity Mach number is well below a value of 0.3. Hence, and the flow is
Aerodynamic studies of ground vehicles are interested in pressure incompressible resulting in forces and moments to be independent of
and viscous generated forces and moments acting on a vehicle and Mach number and results are only dependent on the Reynolds number
the flow field in close proximity to the vehicle. The investigations similarity parameter. In some cases, local flow phenomena can cause
take many forms directed at a comparison of data sets, such as higher flow speeds, and a more detailed analysis is necessary.
between types of methods for the same test article or various Therefore, the Mach and Strouhal numbers as well as the Roughness
geometric scales of a test article and may employ operational testing ratio will also be introduced in this chapter.
(OT), wind tunnel (WT) testing or computational fluid dynamic
(CFD) analysis. Regardless of the approach and method the objective 2.2. Reynolds Number
is to obtain aerodynamic data that represents a full-scale vehicle in
The significance of Reynolds number differs greatly between an
operation. The ability to use one data set to represent another data set
aerodynamically shaped (streamline) body, such as an aircraft, and a
requires that the two test articles have dynamic similarity. Dynamic
bluff body such as a commercial ground vehicle. An aerodynamic
(i.e. Flow) similarity is present when the two test articles have the
body aligned with the flow will have minimal flow-separation and
same length scale ratio, time scale ratio, and force scale (or mass
low drag that is primarily a result of skin friction drag. For this class
scale) ratio. The first and primary requirement for dynamic similarity
of body the importance of Reynolds number is in defining the
is geometric similarity in which all test article dimensions have the
boundary layer state and thereby the dominating viscous drag term.
same linear scale ratio, including the surface roughness ratio (e/L). In
For a bluff body, pressure drag dominates and significant levels of
addition, geometric similarity requires surface angles, flow directions,
flow separation are present where the degree of flow separation is
and relationship of each test article to the physical surroundings to be
dictated by surface geometry and the state of the boundary layer, both
the same. The combination of the same length scale ratio and time
of which are dependent on the Reynolds number.
scale ratio is the definition of kinematic similarity and dynamic
similarity exists when force and pressure coefficients are identical in
Reynolds number is the primary similarity parameter of interest for
concert with kinematic similarity.
aerodynamic studies of commercial ground vehicles. The parameter
is defined as the ratio of the fluid inertia and the shear stress, as a
The most common approach to mathematically express the concept of
result of viscosity, as expressed in Eq. 2.1 [68, 75].
similarity is through dimensional analysis using some form of the
Buckingham Pi theorem [137]. Using a minimum of independent
variables uses dimensional analysis to express flow similarity with
(2.1)
dimensionless parameters. In aerodynamics the number of independent
parameters is obtained based on the requirement of dimensional The dimensionaless Reynolds number is a combination of physical
homogeneity for any equation expressing a valid relationship among parameters viscosity (μ), velocity (U), length (L) and density (ρ),
physical variables. The fundamental dimensions for aerodynamic which are represented by fundamental dimensions length (L),
studies are; mass (m), length (L), time (t) and temperature (T). mass (M) and time (t) as expressed in the following four
dimensional relationships.
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 601

1. Viscosity (μ) ∼ M/(Lt) Aerodynamic similarity is a foundation of sub-scale studies and requires
2. Velocity (U) ∼ L/t the matching of Reynolds number, boundary layer, on and off-body flow
and geometric details for the test article compared to the operational
3. Length (L) ∼ L
conditions for the full-scale vehicle. In practice this is extremely
4. Density (ρ) ∼ M/ L3
difficult for computational studies as well as experimental testing, see
Chapters 4 and 5 respectively. Although the Reynolds number value can
These parameters are used to dimensionally define the inertial forces
be matched in computational studies an accurate simulation of the
and the viscous forces.
relevant boundary layer and flow features can only be obtained if the
Reynolds number dependent boundary layer characteristics and flow
The inertial forces per unit area ∼ (mass flow rate per unit area) x
separation features on the full-scale reference vehicle at operational
(fluid velocity) = ρUU = ρU2
conditions are implemented in the computational model. Coastdown
testing has a different challenge in which the width based Reynolds
The viscous forces per unit area ∼ τ = μ (du/dy) = μU / L
number will typically vary from greater than 5 million to less than 1
million during a test run. The large variation in Reynolds number may
Taking the ratio of the two forces we can derive the nondimensional
generate significant changes in separation characteristics and drag
expression for Reynolds number, see Eq. 2.2.
values resulting in an increase in data uncertainty. An overarching
challenge in defining the Reynolds number sensitivity of a complete
vehicle is accounting for the differences in Reynolds number sensitivity
(2.2) for the various components of the vehicle. In addition to the variation in
Reynolds number sensitivity for each vehicle component, bluff body
For commercial vehicles the vehicle width (w) is typically used as the
vehicles have a significant Reynolds number sensitivity with changes in
reference length (L) resulting in nominal operational Rew values of
yaw angle. See Chapters 3, 4, 5 and 6 for additional information.
0.8, 2.4 and 4.8 million for vehicle speeds of 10, 30, and 60mph,
respectively. However the geometric complexity of this class of
vehicle results in significant local Reynolds number effects that are 2.3. Principal Elements of Aerodynamics
better represented by a component based reference length. Typical The fundamental principles from which the equations used to model
examples are the trailer front side edge radius, diameter of tire/wheel low speed aerodynamic flows are;
assembly and chord length of the cab visor.
1. Conservation of mass;
Vehicle Reynolds number is typically calculated based on a 2. Newton's second law (Force = mass x acceleration);
fundamental geometric length such as width (w) of the vehicle. An
3. Energy is conserved.
equivalent reference length is the square root of the reference vehicle
area (A). Component Reynolds number values shall be calculated
The equations expressing these three principles provide the
based on the linear dimension of the component in the streamwise
relationships among various quantities (such as density, velocity,
direction. Local Reynolds number values shall be based on the local
pressure, rate of strain, internal energy, and viscosity) as they vary in
radius of curvature (r) or the linear dimension of the region in the
space and time. Additional discussion on this topic can be found in
streamwise direction (l). All Reynolds number values shall be based
references [3, 14, 15, 38, 40, 84, 136, 138, 139, and 140].
on the operating conditions of the reference vehicle and the test/
analysis conditions of the model/geometry under investigation.
In considering low speed flows it is common to adopt the assumption
that the density is constant. In cases in which density is nearly
Calculation of Reynolds number based on model/vehicle width (w) is
constant, there are many situations in which the temperature variation
shown in Eq. 2.3.
is negligible. Such problems are entirely mechanical without any
thermodynamic phenomena. The force and moment coefficients in
such cases will, of course, not be dependent on Mach number.
(2.3)

Calculation of Reynolds number based on square root of the model/ However, in pursuit of flow similarity local compressibility effects
vehicle reference area (A) is shown in Eq. 2.4. can occur for configurations with small edge radii, when testing at
incompressible speeds. In such cases, both compressibility and
Reynolds number can influence flow separation onset and effects, and
at times one effect can mask the other. Experimental Reynolds
(2.4)
number effects can also be determined when the free stream velocity
Calculation of Reynolds number based on model/vehicle local radii is held constant and the geometry or fluid density is varied [3].
of curvature (r) of interest is shown in Eq. 2.5.
In addition to Reynolds number there are three other non-dimensional
parameters that are considered in subsonic aerodynamic studies. These
parameters are; Mach number (Ma), Strouhal number (St) and
(2.5)
Roughness ratio (Rr). Mach number is defined as the ratio of the
freestream velocity (U) to the speed of sound (c), based on freestream
602 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

flow, see Eq. 2.6. For Mach number values less than 0.3 the flow is To relate the importance of Reynolds number to the aerodynamic
incompressible and the density is constant. Equation 2.7 is the definition loads Anderson [3] presents a dimensional analysis using the
of the Strouhal number (St) that is used to characterize flow oscillations. Buckingham Pi (P) theorem [135] that shows force and moment
The Strouhal number is defined as the ratio of the product of the flow coefficients are only a function of Reynolds number and Mach
oscillation frequency (ω) and reference length (L) to the freestream number for a given orientation of the test article relative to the flow
velocity (U). Additional discussion is presented in sub-section 2.5. The direction. A brief summary of this analysis follows;
final parameter is the Roughness Ratio (Rr) defined in Eq. 2.8. This
parameter is used to guide the modeling of the surface roughness height The aerodynamic force on a body can be expressed as a function of
(ε) for a sub-scale model compared to that for a full-scale vehicle. physical variables; force (F), density (ρ), velocity (U), length (L),
viscosity (μ), and speed of sound (c) as shown in equation 2.14.

(2.6)
(2.14)

(2.7) Following the Buckingham theorem, the number of dimensionless


quantities needed to fully describe a physical system is the difference
of the number of independent variables with dimension, here the six
(2.8) present in equation 2.14, and the number of physical dimensions
involved, here three (length L, mass M, time t). Hence, equation 2.14 is
2.4. Aerodynamic Loads re-expressed using 6-3=3 nondimensional Π relationships in equation
2.15. The three Π relationships are defined in equations 2.16, 2.17 and
Although the primary focus of aerodynamics is to understand the
2.18 in which the right side of each equation is dimensionless. Solving
forces and moments, these values are built upon more fundamental
for the coefficient of each physical variable in equations 2.15, 2.16 and
elements such as dynamic pressure (Q) defined in Eq. 2.9. Dynamic
2.17 allows further development of the equations.
pressure is used to obtain the non-dimensional pressure coefficient,
skin friction coefficient, force coefficients and moment coefficients
defined in equations 2.10, 2.11, 2.12, and 2.13 respectively. For
incompressible flow the dynamic pressure is exactly the difference (2.15)
between total and static pressure.

(2.16)

(2.9)

(2.17)
The loads acting on a body immersed in a flow are produced by
pressure forces acting normal to the surface and tangential stress (skin
friction) over the surface, see equations 2.10 and 2.11 respectively.
(2.18)

Substituting equations 2.16, 2.17 and 2.18 into equation 2.15 and
(2.10) reorganizing results in equation 2.19.

(2.11) (2.19)

When integrated, these stresses give rise to the resultant force and For the subsonic operational conditions of commercial vehicles we
moment load components, which are expressed in non-dimensional know that Mach number effects are not present resulting in the
form by means of force and moment coefficients, defined as follows: aerodynamic forces and moments on a commercial vehicle are only a
function of Reynolds number. This fact magnifies the importance of
flow similarity in the aerodynamic study of commercial vehicles.
(2.12)
2.5. Time Dependence
Aerodynamic loads have, in general, mean and time-varying
(2.13)
components. The fluctuating loads may be significant not only when
where F and M are the components in the direction of the resultant the upstream flow is time-dependent but also when the wake
force and moment, respectively, acting on the body, U is the produced by the body itself has more or less regular fluctuations. For
undisturbed upstream flow velocity, A is the reference area typically steady upstream flow an aerodynamic body is characterized by a
defined as the projected frontal area, ρ is the density of the flow steady wake and load, whereas for bluff body commercial vehicles
stream and L a reference length.
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 603

the opposite is true [141]. Specific examples related to commercial relative growth in boundary layer thickness as it transitions from
vehicles involve trailing wakes, rotating systems, and multiple bodies laminar to turbulent. The boundary layer state and the transitioning of
such as towed components. the boundary layer between states is a function of Reynolds number.

When the boundary layer separates prematurely on a bluff body, the For the case of a specific surface geometry, such as a flat plate,
phenomenon of separation is associated with the formation of Reynolds number may be used to infer the boundary layer state as
vortices and a large energy loss in the wide wake. At moderate either laminar or turbulent, referenced to the stagnation point. Flat
Reynolds numbers the wake is dominated by a periodic train of plate transition from laminar to turbulent flow typically occurs for
alternating vortices (see Figure 2.1), known as the von Kármán Reynolds number values based on local length between 1 and 2
vortex street [142], at moderate Reynolds numbers. million but may be as low as 0.5 million [117]. A sketch comparing
the features of a flat plate boundary layer with increasing Reynolds
number are depicted in figure 2.3. The sketch indicates the forward
movement of the transition region with increasing Reynolds number.
Photographs of streamwise sectional cuts of laminar, transitional and
turbulent boundary layers are shown in figure 2.4 which highlight the
increased mixing as the boundary layer transitions to turbulent flow.

Figure 2.1. von Kármán vortex street from a circular cylinder (flow is left to
right) [142].

A vortex street is present when boundary layers separate from the


opposing sides of a body roll up in the wake region producing large
vortices that separate from the body and move downstream as free
vortices. The mechanism of shedding is found to be the interaction
between the boundary layers from either side of the body. The
predominant frequency of shedding is a narrow band of values such
that when it is presented in a non-dimensional form, is called the
a. Tangential velocity in laminar boundary layer made visible by tellurium
Strouhal number. The presence of oscillatory vortex shedding will
seeding [142].
impact the aerodynamic forces on the body. In certain regions of
Reynolds number, this shedding process is inhibited and a coherent
set of vortices is not produced although a random shedding process of
smaller scale still occurs.

2.6. Boundary Layer Flows


In 1904 Prandtl [83] introduced the concept of a boundary layer, as a
thin region close to the solid surface where the effects of viscosity are
felt, and a large velocity gradient exists normal to the surface.

The boundary layer is often referred to as a shear flow. A boundary


layer begins to form at the stagnation point on a body and increases
in thickness with distance along the surface, except in regions of high
acceleration where its thickness can be reduced.

Shown in figure 2.2a is a photographic image taken by Wortman


[142] of the velocity profile in a laminar boundary layer and depicted
in figure 2.2b are measured velocity profiles for laminar, transitional
b. Velocity distribution
and turbulent boundary layers. The thickness of the velocity boundary
layer is defined as the distance from the solid body at which the Figure 2.2. Boundary layer velocity profile over a flat plate (flow is from left
viscous flow velocity (u) is 99 percent of the freestream velocity (U). to right).
Note the difference in boundary layer profile, thickness and the
604 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

Transition of the boundary layer from laminar to turbulent depends with decreasing Reynolds number. For example, a flat plate analysis
on Reynolds number and is influenced by parameters such as wall indicates that at 1.0 million, representing a full scale vehicle
surface roughness, surface curvature, freestream turbulence, and the operating at low speed or testing a sub-scale model, laminar boundary
pressure gradient along the surface. These features may promote or layer flows may be present over a significant portion of the front of
inhibit boundary layer transition. Additionally, the specific flow the vehicle and the development of a fully turbulent boundary layer
features within a boundary layer will vary with Re, surface will occur significantly downstream compared to the full-scale
roughness, crosswind, temperature, freestream turbulence as well as vehicle operating at highway speeds. The result is a change in
other factors. Turbulent boundary layers are characterized by an separation characteristics and vehicle drag. Additional details can be
increased value of the tangential stresses at the surface producing found in Schlichting [117].
larger friction forces on the bodies.

2.6.1. Laminar Boundary Layer


All boundary layers begin as laminar (see figures 2.2, 2.3 and 2.4) in
which the streamwise velocity changes uniformly as one moves away
from the body and all exchange of mass or momentum takes place
only between adjacent layers within the boundary layer. Laminar
boundary layers will always be present but the ability to persist along
a body require; that external disturbances do not excite resonances
within the layer, a favorable external pressure gradient as well as a
sufficiently smooth body surface.

Approximations of the laminar boundary layer thickness,


displacement thickness, skin friction coefficient and integrated
average skin friction drag coefficient for a flat plate of length l are
Figure 2.3. Effect of Reynolds number on boundary layer flow. provided in equations 2.20, 2.21, 2.22 and 2.23, respectively. The
variation of these parameters with Reynolds number will be discussed
in sub-section 2.6.3.

(2.20)

(2.21)

(2.22)

(2.23)

2.6.2. Turbulent Boundary Layer


A turbulent boundary (see figure 2.3 and 2.4) layer exist downstream
of the laminar to turbulent transition region and is characterized by
rapid fluctuations in velocity and pressure due to the eddying motion
with mixing across several layers resulting in an exchange of mass,
Figure 2.4. Flat plate boundary layer flow structure [142].
momentum and energy on a much larger scale compared to a laminar
boundary layer.
As with all boundary layers those found on operating commercial
road vehicles or their components begin as laminar and transition to Approximations of the turbulent boundary layer thickness,
turbulent as the flow moves aft along the surface of the vehicle. At displacement thickness, skin friction coefficient and integrated
width-based Reynolds number for a commercial vehicle at highway average skin friction coefficient for a flat plate are provided in
speeds the boundary layer on the forward portion of the vehicle or equations 2.24, 2.25, 2.26 and 2.27, respectively. The variation of
vehicle component, such as the tractor of a tractor-trailer, will have these parameters with Reynolds number and surface roughness
small but measurable percentages of both laminar and turbulent effects will be discussed in sub-section 2.6.3.
boundary layers, however, the boundary layer on the trailer of a
combination vehicle will be predominately turbulent. The extent of
laminar flow and the thickness of the boundary layer will increase (2.24)
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 605

(2.25)

(2.26)

(2.27)

2.6.3. Natural Laminar-Turbulent Transition


Boundary layer transition takes place over a range of critical
Reynolds numbers where the characteristic length in the definition of
Reynolds number is the distance along the surface from the
stagnation point. The transition Reynolds number is affected by many
factors; most important are surface pressure distribution, the surface
roughness and the intensity of free stream turbulence.

The start of transition is characterized by short turbulent bursts with


rapid velocity and pressure fluctuations; see figure 2.5. Further
downstream the duration and frequency of the turbulent signals
increase until the boundary layer is fully turbulent. A boundary layer Figure 2.6. Drag coefficient of laminar, turbulent and transitional boundary
can transition to turbulent through a number of paths depending on the layer on flat plate [143].
conditions such as initial disturbance amplitude and/or surface
roughness. The first stage of the natural transition process consists of The impact of boundary layer state, boundary layer thickness,
the transformation of environmental disturbances into small boundary layer displacement thickness, friction drag and transition
perturbations within the boundary layer. The next stage of the transition Reynolds number on the aerodynamic performance of commercial
process is that of primary mode growth in which the initial disturbances vehicles will change with each design cycle of the vehicle. Vehicles
grow or decay. The specific instabilities that are exhibited depend on designed prior to 1990 have limited aerodynamic contouring. As a
the body geometry and the nature and amplitude of initial disturbances. result pressure drag dominated despite the fact that the friction drag
was also maximized due to a combination of transitional Reynolds
numbers below 106 and the presence of a rough surface covering
large areas of the vehicle. Vehicles from the 1990 to 2000 era have
increased aerodynamic shaping and routinely employ aerodynamic
fairings on the tractor to reduce the pressure drag. These vehicles
have reduced surface roughness as a result the transition Reynolds
number is increased however the fairings increase the wetted area
resulting in friction drag that is equivalent to pre-1990 vehicles.
Recent design studies indicate that the next generation of vehicles
will have significant reductions in pressure drag that will effectively
increase the percent contribution of friction drag to the total drag.
These vehicles will have further increase in the transition Reynolds
number, more extensive presence of laminar flows and thinner
boundary layers. This combination will combine to increase the
Figure 2.5. Photograph of turbulent spot in a transitional boundary layer, flow
sensitivity of the vehicle to local flow separations and roughness
left to right [142].
induced drag. Additional discussion of these design trends will be
A graph of the variation in friction drag for laminar, transitional and presented in Chapters 5 and 6.
turbulent boundary layer flows as a function of Reynolds number is
presented in Figure 2.6. The figure shows flat-plate friction drag 2.6.4. Forced Laminar-Turbulent Transition
coefficients for laminar and turbulent boundary layer smooth-wall
At Reynolds number values below the operational Reynolds number
relations as well as the effect of surface roughness. In the fully rough
of a vehicle, a laminar boundary layer flow is present over a greater
regime, which is plotted to the right of the dashed line in Fig. 2.6, Cf
surface area of the vehicle (or model) compared to that at operational
is independent of the Reynolds number, so that the drag varies as U2
conditions. Relative to a turbulent boundary layer a laminar boundary
is based on Schlichting [117]. The figure also shows the behavior of
layer is more susceptible to flow separation, when subjected to an
the drag coefficient in the transition region 5×105 < Ret < 8×107,
adverse pressure gradient, resulting in flow features and aerodynamic
where the laminar drag at the leading edge is an appreciable fraction
loadings that are not representative of the vehicle in operation. In
of the total drag.
contrast a turbulent boundary layer is able to withstand a more
606 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

intensive adverse pressure gradient without separating. Matching 2. Boundary layer transition roughness elements should be
full-scale boundary layer features such as the location of boundary randomly distributed or have an irregular shape.
layer transition, boundary layer thickness and location of separation 3. The roughness should be a minimum width strip that is aligned
requires a detailed understanding of boundary layer, flow control perpendicular to the surface streamlines.
concepts as well as the boundary layer characteristics at full scale 4. Boundary layer transition roughness shall have a minimum
Reynolds number. Successful use of forced transition to mimic high roughness Reynolds number (Rek) of 600.
Reynolds number boundary layer flow requires an understanding of
5. Forced transition application should vary with vehicle shape,
the desired boundary layer conditions as well as the expected
vehicle type, operational conditions, and test objectives.
aerodynamic characteristics at higher Reynolds number conditions.
6. Forced transition effectiveness shall be verified experimentally.
The aerodynamics of a vehicle can change significantly with changes
in Reynolds number and/or changes in boundary layer state. It is
known that the critical and transcritical Reynolds number can occur
at lower values by introducing turbulence structures into the
boundary layer to promote early laminar to turbulent transition.
Methods used to force transition include roughness elements, vortex
generators and trip wires placed on the surface of the body, heating of
the surface and oscillatory blowing of the boundary layer.

As shown in figure 2.6 surface roughness will cause the boundary


layer to transition from laminar to turbulent at a lower Reynolds
number compared to a smooth surface. The result is an increase in
friction and total drag, compared with the smooth surface value, but
under certain conditions the reverse happens as illustrated in Figure
1.3. At certain Reynolds numbers the laminar boundary layer
separates upstream resulting in a wide wake and a large pressure
drag. However, if the boundary layer transitions to a turbulent state Figure 2.7. Details of distributed roughness criteria for forced boundary layer
before separation occurs then it is less susceptible to separation and transition [117].
remains attached further round the body surface compared to the
laminar boundary layer. This results in a reduction in wake width,
2.6.5. Boundary Layer State Assessment
pressure drag and total drag, relative to the smooth surface values.
To understand the effect of Reynolds number on the aerodynamics of a
vehicle/model requires an understanding of the boundary layer
In use since 1940, surface roughness is a generally accepted approach
characteristics over the complete vehicle/model, [44, 118, 136, 139,
where roughness elements take many forms from individual grit
142, 145, 149, and 150]. Of particular importance is the location of
elements, shaped and positioned 2D structures, tape strips, vortex
transition. There are numerous intrusive and nonintrusive techniques
generators, and sandpaper, see Barlow [14], Metz [44], Loftin [144],
that may be used. These methods vary from a boundary layer survey
Dryden [145], Klebanoff [146], Braslow [147], [117), and Von
rake/probe to optical techniques such as a laser Doppler velocimetry
Doenhoff [148]. Most surface roughness techniques have been
(LDV) to determine the velocity profile in the boundary layer. Hot
developed to support aircraft related subscale wind tunnel testing in
films or floating element balances may be placed on the model surface,
which the work of Von Doenhoff [148] and Braslow [116] are most
or pressure sensors may be placed flush with the surface or behind a
significant, see figure 2.7. The objective is to force transition with
surface pressure tap, in order to measure the fluctuations due to
roughness elements such that the drag of the roughness elements
turbulence. These three methods have the advantage that they are
produces a negligible amount of roughness drag (i.e. roughness drag is
essentially non-intrusive and allow simultaneous measurements. A
the pressure drag of the roughness elements). For the study of low drag
disadvantage is that they do not give useful information in regions of
aerodynamic shapes roughness drag can be a significant portion of the
separated flow. Each measurement point requires its own transducer
drag, however for ground vehicles roughness drag is not a significant
and signal conditioning equipment. Oil film interferometry and a hot
portion of the vehicle drag. For subsonic boundary layers the roughness
wire may be placed on the surface of the body. Stanton gauge, Preston
size requirements are reasonably similar for 2D and 3D flows.
tubes and survey rake probes may be traversed along the surface. These
techniques allow for a high density of measurement points. However
Forced transition should be used to simulate full-scale Reynolds
there are errors associated with the intrusive nature of a physical probe.
number flows, match boundary layer thickness, and inhibit separation
of the boundary layer. See Chapter 5 for additional guidelines. The
When evaluating a complex vehicle/model consisting of multiple
general roughness application criteria are;
components the location of transition and the boundary layer state
should be determined at various locations on the vehicle/model as
1. Boundary layer transition roughness elements shall have a
well as on each component. When comparing test data for different
height that is less than the thickness of the laminar boundary
Reynolds number that have been obtained from the same or different
layer at the location applied.
test methods the location of transition and the boundary layer state at
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 607

various locations should be determined for each data set. For


example, for wind tunnel tests of a vehicle/model a boundary layer
survey should be used to obtain the location of transition and the
boundary layer profile immediately upstream of the trailing edge of
each vehicle/model component.

2.6.6. Boundary Layer (Flow) Separation


Experimental methods used to evaluate vehicle aerodynamic
performance routinely involve test conditions at Reynolds number Figure 2.9. Representative change in boundary layer profile for the onset of
significantly less than that seen by the operational vehicle. separation.
Understanding and controlling boundary layer separation for
Reynolds number at and below the operational Reynolds number is Flow visualization images showing the impact of favorable and
critical to testing, analysis and vehicle design efforts. The tendency of unfavorable static pressure gradients on an axisymmetric body with a
a boundary layer to separate is dependent not only the Reynolds laminar boundary layer are presented in figure 2.10. Figure 2.10a show
number and boundary layer state but also the surface pressure flow visualization for a bluff face body with moderate curvature at the
gradient which is dictated by the shape of the vehicle. The influence shoulder and figure 2.10b show a body with increased nose curvature.
of a pressure gradient on the boundary layer profile is shown in figure The static pressure distribution on bluff body of figure 2.10a would
2.8a, 2.8b and 2.8c respectively. As shown in figure 2.8b a negative have a pronounced negative gradient followed by a severe positive
static pressure gradient (dP/dx < 0) is described as favorable because gradient as the flow transitions from the front face and around the
the boundary layer is stabilized by the increase in freestream velocity. curved shoulder of the body. The severe positive gradient (i.e.
This type of gradient would be observed in a region of smooth unfavorable) is the mechanism for the observed separation. In contrast
convex curvature. In contrast a positive static pressure gradient (dP/ the static pressure distribution on the body shown in figure 2.10b would
dx > 0) is unfavorable as it produces a deficit in the boundary layer have a gradual positive gradient (i.e. favorable) as the flow moves aft
velocity profile eventually leading to reverse flow and separation as from the stagnation point at the forward most point on the body.
depicted in figure 2.8d. Possible locations on a body where an
unfavorable static pressure gradient would occur are the transition The mixing process in a turbulent boundary layer causes an effective
region from a convex curvature region to a flat surface or within a shear stress which allows a turbulent boundary layer to progress
region of concave curvature. further against an adverse/unfavorable pressure gradient and
separates at a point further along a surface, compared to a laminar
A schematic depicting the change in boundary layer flow during the boundary layer under the same conditions. A comparison of laminar
separation process is shown in figure 2.9. Note that the flow in the and turbulent flow separation over the same geometry is depicted in
figure is from left to right. Boundary layer separation, and thus flow figure 2.11. The photographs depict the flow over a curved surface
separation, occurs when the portion of the boundary layer closest to with an adverse pressure gradient resulting in a high momentum
the wall reverses in flow direction as shown in the third from left deficit. In the case where the boundary layer is laminar, insufficient
boundary layer profile in figure 2.9. The separation point is located momentum exchange occurs, the flow is unable to adjust to the
between the forward and backward flow, where the shear stress is increasing pressure and separates from the surface. In the case of
zero. The overall boundary layer initially thickens at the separation turbulent flow, the increased transport of momentum from the
point and is then forced off the surface by the reversed flow at the free-stream to the wall increases the streamwise momentum in the
surface. Flow separation often results in increased pressure drag. boundary layer. This allows the flow to overcome the adverse
pressure gradient and will delay separation to a point further
downstream.

Figure 2.10. Laminar boundary layer flow over two bluff face body of
revolution [142].

Figure 2.8. Representative flat plate boundary layer velocity profiles.


608 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

Figure 2.12. Effect of free stream turbulence and Reynolds number on the
Figure 2.11. Example of laminar (upper photo) and turbulent (lower photo) drag of a sphere [104].
separation over a curved surface [142].
The aerodynamics and performance of a commercial vehicle can be
The presence of a body in relative motion to a fluid will cause the affected in various ways by freestream turbulence in both the wind
flow passing over the body to accelerate and decelerate resulting in tunnel testing as well as for a full-scale vehicle in service or subjected
favorable and unfavorable pressure gradients. When the body is thin, to on-road or on-track testing [54, 153, 154, 155]. The impact of
the pressure gradients are weak and the flow remains attached. For a turbulence on the aerodynamics of a vehicle or model is a function of
thicker and less aerodynamic shaped body, adverse pressure gradients turbulence intensity and length scale and will vary with test speed,
cause strong flow decelerations resulting in flow separation, yaw angle, vehicle shape, transition Reynolds number and boundary
circulation, and vortex shedding. For a commercial vehicle the rate of layer state. The interaction between turbulence and a body is unique
change in body thickness is exaggerated resulting in large pressure and does not lend itself to general trends in aerodynamic forces and
gradients and susceptibility to flow separation and a drag level that is moments [54, 154, 155].
sensitive to Reynolds number. For the range of operational conditions
of a commercial vehicle flow separations that occur at the front of the Operational vehicles will experience large variations in turbulence
vehicle in close proximity to a stagnation point or from interactions that can generate a broad range of effects [152, 153]. Turbulence
with a sharp edge or corner will typically not be sensitive to Reynolds scales that exceed the size of a vehicle typically have a negligible
number. If the geometry at the front of the vehicles is void of sharp effect on vehicle loading other than changing flow direction. When
edges and abrupt geometry changes then flow separations may be turbulence size is equivalent to vehicle size it can interact with
sensitive to Reynolds number. However flow separation at the trailing vehicle local surface loadings that may impact flow separation and
edge of the vehicle will be sensitive to Reynolds number. total drag. Small-scale turbulence levels have the ability to alter local
boundary layer characteristics that will impact flow separation,
reattachment and wake flows [135].
2.7. Free Stream Turbulence Effects
Commercial vehicles in operation on the open road experience
turbulence [151] from a variety of sources with average levels 2.8. Summary Findings
between two percent and five percent and turbulent length scales that A discussion of Reynolds number modeling and simulation criteria
exceed the width of a typical vehicle [151, 152]. In contrast the free for fundamental aerodynamic and fluid dynamic concepts that impact
stream turbulence intensity and length scale in a typical wind tunnel the aerodynamic characteristics and performance of a commercial
are one to two orders of magnitude smaller [31, 50, 54, 58, 60, 96, vehicle are presented. The concepts of dimensional analysis and flow
104]. The effect of turbulence can be best demonstrated with a wind similarity are presented for various fluid dynamic topic areas. These
tunnel test of a smooth shaped aerodynamic body in which both concepts are applied to show that the aerodynamics of commercial
laminar and turbulent boundary layers are present. Figure 2.12 from ground vehicles is only dependent on the Reynolds number similarity
Diehl [104] show the effect of free stream turbulence on a smooth parameter. The critical subsonic non-dimensional parameters as well
sphere is to promote boundary layer transition resulting in drag as the non-dimensional aerodynamic load parameters are developed
reduction to occur at a lower Reynolds number. Diehl states that these using the dimensional analysis process.
results are not exact but are representative in that similar test have
shown that sphere size and turbulence length scale will alter these A significant portion of the Chapter focused on boundary layer flows,
results. This and other similar data have provided the motivation to specifically the impact of boundary layer state, boundary layer
minimize the turbulence in all current wind tunnels. thickness, boundary layer displacement thickness, friction drag and
transition Reynolds number on the aerodynamic performance of
However the desire to minimize turbulence intensity and scale is at commercial vehicles. The boundary layer discussion reviewed the
odds with the ability to model the flow experienced by a commercial scaling and simulation challenges related to sub-scale wind tunnel
vehicle in operation [152]. testing and correlating wind tunnel and full-scale fuel economy or
road load tests. A discussion of transition reviewed both free and
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 609

forced transition effects and methods. The influence of free stream 146. Klebanoff, P., Schubauer, G. and Tidstrom, K., “Measurements of the
Effect of Two-Dimensional and Three-Dimensional Roughness Elements
turbulence on transition and drag was also reviewed and highlighted on Boundary Layer Transition,” Journal of the Aeronautical Sciences
the challenge facing sub scale aerodynamic testing of current and 22(11): 803-804, 1955.
future vehicle designs.2.9. 147. Braslow, A. and Knox, E., “Simplified Method for Determination of
Critical Height of Distributed Roughness Particles for Boundary-Layer
Transition at Mach Numbers From 0 to 5,” NACA TN 4363, 1958.
148. Von Doenhoff, A. and Horton, E., “A Low-Speed Experimental
2.9. References Investigation of the Effect of a Sandpaper Type of Roughness on
136. Shames, I., “Mechanics of Fluids,” 2nd Ed., McGraw-Hill Inc., 1994 Boundary-Layer Transition,” NACA TN 3858, 1956.
137. Buckingham, E., “On Physically Similar Systems: Illustration of the Use 149. Squire L., “The Motion Of A Thin Oil Sheet Under the Boundary Layer
of Dimensional Equations,” Physical Review 4(4):345-376, 1914. See on a Body,” AGARDOgraph no. AG-70: 7-28, 1962.
also:“Model Experiments and the Forms of Empirical Equations,” Trans.
ASME, 37:263-288, 1915. 150. Crook, A., “Skin Friction Estimation at High Reynolds Numbers and
Reynolds Number Effects for Transport Aircraft,” Center for Turbulence
138. Baker, W., Westine, P., and Dodge, F., “Similarity Methods in Research, Annual Research Briefs, pp. 427-438, 2002.
Engineering Dynamics - Theory and Practice of Scale Modeling,”
Revised Edition, Elsevier, 1991 151. Engineering Sciences Data Unit, “Characteristics of Atmospheric
Turbulence Near the Ground Part II: Single Point Data for Strong Winds
139. Karamcheti, K., “Principles of Ideal Fluid Aerodynamics,” John Wiley (Neutral Atmosphere),” ESDU Data Item 85020, ISBN: 0-85679-526-7,
& Sons, 1966, Chapter 1 2001.
140. Sedov, L., “Similarity and Dimensional Methods in Mechanics,” 10th 152. McAuliffe, B., Belluz, L., and Belzile, M., “Measurement of the On-
Edition, CRC Press, 1993. Road Turbulence Environment Experienced by Heavy Duty Vehicles,”
141. Landahl, M. and Mollo-Christensen, E., “Turbulence and Random SAE Int. J. Commer. Veh. 7(2):685-702, 2014, doi:10.4271/2014-01-
Processes in Fluid Mechanics,” Second Edition, Cambridge University 2451.
Press, ISBN: 9780521422130, 1992. 153. Bearman, P. W. “Review - Bluff Body Flows Applicable to Vehicle
142. Van Dyke, M. “An Album of Fluid Motion,” The Parabolic Press, Aerodynamics,” ASME Journal of Fluids Engineering 102(3):265-274,
Stanford, ISBN: 0-915760-02-9, 1982 1980.
143. White, F. “Fluid Mechanics,” 7th Ed, McGraw Hill, ISBN: 0077422414, 154. Morel, T., “Aerodynamic Drag of Bluff Shapes: Characteristics of Hatch
2009 Back Cars,” General Motors Research Laboratories Report GMR-2581,
144. Loftin, L. Jr., “Effects of Specific Types of Surface Roughness on 1977.
Boundary-Layer Transition,” NACA WR L-48, 1946. 155. Martin, L. J., “The Effect of Turbulence on the Flow Around a Cube,”
145. Dryden, H., “Review of Published Data on the Effect of Roughness on M.Sc. dissertation, Department of Aeronautics, Imperial College, 1977.
Transition From laminar to Turbulent Flow,” Journal of the Aeronautical
Sciences 20(7): 477-482, 1953.
610 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

CHAPTER 3. BLUFF BODY AERODYNAMICS turbulent flow and the drag level becomes invariant with further
increase in Reynolds number.
A bluff body is a class of aerodynamic shapes that includes all
commercial vehicles. Bluff body vehicles have blunt fronts with 5. Transcritical Reynolds number Range may exhibit small
truncated bases and their aerodynamics are chracterized by variations in drag due to local changes in boundary layer state,
appreciable levels of flow separation and pressure drag levels that are flow separations, and flow reattachment resulting in changes in
significantly larger than friction drag. Reynolds number is used in the ratio of viscous to pressure drag.
bluff body aerodynamic studies to characterize on-body and off-body
flow features such as the vehicle boundary layer state, boundary layer
thickness, flow separation, flow expansion, shear layer stability and
flow reattachment. Each of these fluid dynamic features plays an
important role in guiding experimental as well as computational
studies of the drag force acting on a commercial vehicle. Specific
aerodynamic design and analysis areas of interest for commercial
vehicles are wind tunnel testing, coastdown testing, computational
design, advanced surface shaping, flow control technology and time
dependent atmospheric effects. The Reynolds number and boundary
layer dependent aerodynamic trends associated with specific bluff
body geometrical features found on commercial vehicles will be
discussed in this Chapter. The supporting bluff body aerodynamic
data and information can be found in references 1, 5, 10, 12, 13, 14,
17, 18, 20, 21, 23, 24, 25, 26, 27, 28, 30, 36, 37, 38, 39, 40, 42, 43,
45, 46, 48, 49, 51, 52, 53, 54, 56, 57, 60, 65, 68, and 154, 155, 156,
157, 158, 159, 160, 161, 162, 163, 164, 165, 166, 167, 168, 169, 170,
171, 172, 173, 174, 175, 176, 177, 178, 179, 180, 181, 182, 183, 184,
185, 186, 187, 188, 189, 190, 191, 192, 193, 194, 195, 196, 197, 198,
199, 200, 201, 202, 203, 204, 205, 206, 207, 208, 209, 210, 211, 212, Figure 3.1. Representation of the variation in CD with Reynolds number for a
213, 214, 215, 216, 217, 218, 219, 220, 221, 222, 223, 224, 225, 226, sphere and a typical bluff vehicle.
227, 228, 229, 230.
Both curves reflect the variation in drag with increasing Reynolds
number starting from a condition in which the boundary layer is fully
3.1. Terminology/State Conditions laminar to the condition where the boundary layer is fully turbulent. The
The first step in understanding the impact of Reynolds number on the curves represent the variation in drag over a Reynolds number range of
aerodynamics and specifically the drag of a vehicle is to define the 105 to 107 which includes the operational range of Reynolds numbers
range of Reynolds numbers and the corresponding variation in drag for medium and heavy commercial vehicles of 0.5 to 6.0 million.
that reflects the operational as well as test/analysis conditions of these However the vertical scales for the two curves are not equivalent, as the
vehicles. This information establishes the boundaries for aerodynamic sphere will experience a drag reduction of 70 to 80 percent while the
study of commercial vehicles. A graphical representation of the bluff body curve reflects a drag reduction on the order of 10 to 20
variation in drag with Reynolds number for current-model commercial percent. Experimental tests by Dryden [31], Jacobs [98], Hoerner [38,
vehicles and a sphere is presented in figure 3.1. The sphere is 99], Diehl [104] and Platt [201] serve as the basis for the sphere curve.
presented as a reference and is selected because it is the most The bluff-body vehicle curve is a composite of wind tunnel data for
fundamental of three-dimensional bluff body shapes, as discussed generic shapes and representative vehicles from Wood [2], Cooper [26],
previously in Chapters 1 and 2. The variation in drag, depicted in each Flynn [32], Sovran [54], Storms [55], Leuschen [65], McArthur [66],
curve, is a result of changes in the boundary-layer state, flow Gross [124], Turner [128], White [143], Kirsch [174], Ortega [181],
separation and wake flow. To guide the remainder of the discussion Kettinger [193], Van Raemdonck [194], Littlewood [195], Hjelm [196],
each curve is classified into five regions where each region represents Newnham [197], Heft [198], and Storms [199].
a change in the fundamental flow characteristics as defined by Roshko
[49]. A description of each Reynolds number region is defined below.
3.2. Fundamental Aerodynamics
1. Subcritical Reynolds number Range represents the condition in While the curves in figure 3.1 represent the drag variation for two
which the boundary layer and all separations are laminar. shapes that differ greatly in geometric complexity they also reflect the
2. Critical Reynolds number (Recr) represents the start of boundary characteristic trends with Reynolds number that would be observed
layer transition from laminar to turbulent. for current vehicles as well as vehicle components. This observation
is not unique to bluff bodies as noted by Diehl in 1928 [104] in his
3. Transitional Reynolds number Range is where large variations
comments related to Reynolds effects for aircraft.
in drag are present indicating the progression in the extent of the
boundary layer transition from laminar to turbulent.
4. Transcritical Reynolds number (Retr) represent the point at
which a majority of the boundary layer has transitioned to
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 611

“Actually the effect of Reynolds Number is different on each item: value between that for the curved edge rectangle and the oval shaped
wings, fuselage, struts, wires, etc. Furthermore, the effect is not the body. Additionally, we see that the use of airfoil shaped section
same on all wings or on all struts, but varies widely in each group.” (second from bottom) is far superior to a simple oval for components
such as a visor, vertical exhaust, mirror, and landing gear.
Diehl's observation highlights the importance and complexity of
Reynolds number effects and the need to understand these effects for
both the complete vehicle as well as each component.

The drag of a bluff body is driven by the flow separation


characteristics, which are a function of the boundary layer state that
in turn is dictated by Reynolds number and the geometrical features
of the body. The front face and bluff base as well as the length of the
body of a streamwise continuous bluff body will determine the drag,
flow separation and the cross flow sensitivities. Of equal importance
are the detailed geometric features such as edge shape. Shown in
figures 3.2 are plots of drag against Reynolds number for simple 2D
and 3D bluff body shapes with various levels of streamwise
curvature. The data contained in figure 3.2 has been taken from
Hoerner [38], Hucho [40], and White [143] where the drag coefficient
for each curve is based on the body maximum cross sectional area.
The data show that bodies with sharp corners or small radii of
curvature, as depicted in the top two curves, exhibit negligible change
in the flow separation characteristics and drag over a broad Reynolds
number range. In contrast the drag of three-dimensional and
two-dimensional bodies with significant curvature in the streamwise
direction, as shown in the lower five curves, varies significantly with Figure 3.3. Variation in CD with Rel due to edge radius for 2D cylinders.
Delaney [30].
increasing Reynolds number.
Presented in figure 3.3 is data obtained by Delaney [30] that show
variations in CD with Reynolds number for two-dimensional cylinders
with different edge radii. These data are similar as that of Lindsey [43]
and Roshko [49] and may be used to represent the flow about a trailer or
freight box front side and top edges or a region of local curvature on a
tractor. The upper curve in figure 3.3 is for a square cylinder with sharp
edges. For the Reynolds number range shown the flow on the up-stream
facing surface is unable to follow around the sharp corner and separates
resulting in a wake that extends downstream and engulfs the complete
body resulting in a constant drag coefficient value. The transcritical
Reynolds number for this body would occur at a Reynolds number value
that is greater than the range shown on the figure. The transcitical
condition indicates that the flow separation from the up-stream corners
has reattached to the upper and lower surfaces of the body. The two
lower curves represent bodies with increasing edge radii and the data
show that the transnscritical Reynolds number is reduced with increasing
edge radii. The presence of a curved up-stream corner allows the flow to
partially follow around the corner, leading to a smaller separation bubble,
Figure 3.2. Drag coefficients of various cylindrical shapes as a function of Re. which can reattach to the upper and lower surfaces of the body, for the
range of Reynolds numbers shown. Depending on the speed of a
The characteristic trends exhibited by all seven shapes provide insight
full-scale vehicle (as represented by the shaded vertical bar) the local
into the drag levels and Reynolds number sensitivity for various types
edge-based Reynolds number value may be either in the subcritical range
of vehicles and their components. The upper four curves are bluff
(sharp edge, upper curve), transitional range (small edge radius, middle
body shapes and the two lower curves represent aerodynamic
curve) or transcritical range (large edge radius, lower curve). These data
(streamlined) bodies. The oval shaped body located between these two
show that the transcritical Reynolds number can vary over a broad
groups has aerodynamics characteristics of both. For example, the
Reynolds number range for small changes to a vehicle component. These
trend for the sharp edge rectangle (second from top) is representative
data further strengthens the importance of modeling the full-scale
of an early model heavy truck with poor aerodynamic styling while
aerodynamics as well as on-body and off-body flow conditions when
the curved edge rectangle (forth from top) reflects a trend we would
conducting experimental and computational studies at sub-scale or
expect to see for current model commercial vehicles. From these data
off-design condition. See Chapter 5 and 6 for additional guidelines.
we can infer that the lower drag limit for a bluff body vehicle is a
612 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

The data presented in the remaining figures of this Chapter will depict An alternate mechanism to surface roughness is to increase the wind
the Reynolds number sensitivities for commercial vehicles and tunnel free stream turbulence. The increased free stream turbulence
various components comprising a vehicle. structures interact with the boundary layer flow to promote early
transition. Shown in figure 3.5 is a plot from a hatch-back study by
Morel [154]. This study investigated the presence of vortices that
3.2.1. High Reynolds Number Simulation (Forced BL
develop along the side-edges of the slanted afterbody. The data show
Transition)
that free stream turbulence forced early boundary layer transition
The on-body flow, off-body flow and the aerodynamics of a vehicle resulting in a change in both the drag levels and the slant angle at
can change significantly with changes in Reynolds number and/or which flow separation occurs.
boundary layer transition. The Reynolds number for a specific vehicle
will vary with a change in the velocity of the air stream (or vehicle
speed) and/or by a change in the density of the air (fluid) passing over
the vehicle surface. It is known that the critical and transcritical
Reynolds number can be forced to occur at a lower Reynolds number
by promoting boundary layer transition with a specified surface
roughness [44, 99, 144, 145, 146, 147, 148] or by introducing
turbulence into the flow upstream of the body [58, 96, 97, 99, 151,
152, 155, 161, 167, 180, 191, 197, 200, 201]. However, the use of
free stream turbulence to adjust boundary layer transition is complex
and may result in unwanted changes to other flow conditions on the
test article. Introduction of free stream turbulence in a wind tunnel
test to mimic the operational conditions of a vehicle may be difficult
due to differences in turbulent length scales. It is critical for any
aerodynamic investigation that boundary layer transition be
controlled such that the desired outcome is obtained. Representative
results for variations in boundary layer transition and the impact of
transition on the CD, are shown in figures 3.4 and 3.5 for generic Figure 3.5. Effect of change in boundary layer transition on CD for a bluff
vehicle components and in figure 3.6 for a combination vehicle. body with different slant angles [154].

Figure 3.4. Effect of boundary layer transition on ReA dependent CD for 3D


cylinders. Cooper [26].

The chart shown in figure 3.4 is from a report by Cooper [26]. The Figure 3.6. Effect of boundary layer transition on Yaw angle dependent CD for
test program investigated leading edge radius effects on generic bluff two combination vehicles [201].
body models over a broad Reynolds number range for both free
boundary layer transition and the use of grit roughness elements to The two previous examples of forced transition were for simple bluff
force boundary layer transition. Test data is for non-dimensional body shapes at zero yaw angle. Shown in figure 3.6 is a data plot by
radius values (η = r/A0.5) of 0.025, 0.050 and 0.063. The data for η = Cooper [201] for two combination vehicles over a range of yaw
0.063 show that the use of grit reduces the critical and transcritical angles. Results are presented for free transition and three levels of
Reynolds number values by 50 percent. In contrast, the forced free stream turbulence. Compared to the free transition data the
transition data for the relative sharp edge body (η = 0.025) varies forced transition data show significant changes in the CD level and the
little from the free transition data indicating that the sharp edge body variation in CD with yaw angle for each free stream turbulence level.
is insensitive to Reynolds number over the test range. However, at low yaw angels the effect of turbulence is to increase
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 613

drag but as higher yaw angles the drag is reduced when compared to
the low turbulence (free transition) data. This observation is in
contrast to that observed in figure 3.5 and raises concerns about the
use of turbulence to achieve a controlled boundary layer transition.

The data presented in figures 3.4, 3.5 and 3.6 highlights the benefit
and complexity of using forced transition techniques to simulate
higher Reynolds number flow conditions. Successful use of any
forced transition method requires an understanding of the boundary
layer conditions that are to be generated as well as the aerodynamic
characteristics that are expected at this artificially modeled higher
Reynolds number condition. For this reason complex vehicle shapes
with multiple components should make use of forced transition
treatment schemes that are tailored to the location on the vehicle or
component in order to obtain aerodynamic characteristics that
represent those found on the full-scale vehicle at operational
conditions. See Chapter 5 for additional guidelines.

3.2.2. Radius Effect


Aircraft design technology and edge shaping techniques found their
way into ground vehicle design beginning in the 1930s. However
associated with any edge shaping there is an increase in the
complexity of the aerodynamics of the vehicle resulting in greater Figure 3.7. Variation in CD with Rer and edge radius for 3D cylinders [47].
sensitivity of the drag force to both the Reynolds number and the
direction of the onset flow (yaw angle). Edge shaping on commercial
ground vehicles has historically used fixed radius shaping on the side
edge of the tractor, trailer, and freight box. Recent design trends show
an increase in the use of complex three-dimensional shaping where
changes in the local radii of curvature will become the figure of merit
in the evaluation of boundary layer transition and Reynolds number
sensitivity. The importance of this parameter will continue to increase
as the front of trucks and trailers adopt more complex shaping
strategies in the design process. Aerodynamic edge radius effect data
and information can be found in references 10, 14, 18, 30, 35, 43, 47,
51, 61, 158, 175, 180, 181, 184, 194, 202 and 203.

One of the earliest data sets and one that is often cited is the 1930
paper by Pawlowski [47], based on the thesis by Frank Wyszynski,
see figure 3.7. The plot of figure 3.7 is constructed by plotting the
Pawlowski data as a function of Reynolds number based on the edge
Figure 3.8. Variation in CD with Rer and edge radius for 3D cylinders [26].
radius. The plot indicates that the critical Reynolds number (Recr,r)
value for the bodies with an edge radius is less than 0.05 million. For The sample of the edge radius test data of Cooper [26] is presented in
the sharp edged body, the Reynolds number needs to be formed with figures 3.8 and a summary of this data is shown in figure 3.9. The
the front width, as no edge radius is present, leading to critical data of figure 3.8 show a variation in CD with ReA for several blunt
Reynolds numbers greater than 0.25 million. The CD data for bodies edge bluff bodies. As discussed previously for figure 3.4 the edge
with radius values of 0.167, 0.250, and 0.333 are in the transcritical radius data is presented as a non-dimensional radius value (η = r/
Reynolds number range. In contrast the data for the 0.083 radius body A0.5). The data of figure 3.8 show that increasing edge radius reduces
indicate it is in the transitional Reynolds number range. A review of both the Recr,A and Retr,A values proportionally to the change in radius
the full data set suggest that the transcritical Reynolds number (Retr,r), value. Cooper extracted the Recr,r and Recr,A values from the data and
based on edge radius, is greater than 0.125 million. To assist the presents these values as front edge radius design boundaries, see
reader in comparing the edge radius data to width-based Reynolds figure 3.9. The vertical design boundary on the left of the figure show
number values, the radius-based Reynolds number value is converted that the Recr,r value is 0.13 million for moderate edge radius values
to a width based Retr,w value by multiplying the Retr,r value by the typically found on commercial vehicles. On the right side of the
ratio of the test article width to the edge radius value. This conversion figure is the companion Recr,A design boundary. To further evaluate
produces a width based Retr,w of 1.5 million. the design boundaries presented in figure 3.9 several additional data
614 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

sources were analyzed and plotted against the Cooper design


boundaries in figure 3.9. Note, using the scaling process discussed
(3.2)
previously a 6-inch radius front edge on a full-scale trailer (η =
0.063) has a width based Recr,w value of 2.0 million. Vehicles/models
with any local radii of curvature less than 6 percent of the square root
(3.3)
of the reference area (A0.5) should use a minimum width based
transcritical Reynolds number value greater than 2.0 million. It is important to note that Hoerner did not take into account pressure
drag upstream of the base. Hoerner relied upon axisymmetric
streamline bodies with blunt bases in which he assumes the total drag is
comprised of a base pressure drag and a friction forebody drag where
the forebody is the portion of the body upstream of the base. These
drag components allows for the derivation of base drag and describes
the relationship between base drag, forebody drag and total drag.

Hoerner's results have served as a foundation for ground vehicle


research by NASA in in the 1970s and 1980s [48, 186, 204, 205, 206,
207, 208] and generic bluff body studies by Van Raemdonck [194] in
2012. Van Raemdonck's results support the relationship between the
boundary layer and base drag and found acceptable correlation with
Hoerner as modified by Butsko, see equation 3.2. However the basis
of Hoerner's work does not correlate to the drag elements found on a
commercial vehicle. Recognizing this limitation NASA research
re-defined the numerator of equation 3.1 as 0.09 for a lower boundary
Figure 3.9. Comparison of front edge radius data sets with Coopers design and to 0.10 for an upper boundary.
boundaries [26].
To narrow the focus of bluff base flows and drag as they relate to
3.2.3. Base Flows commercial vehicles the following analysis will be limited to boattail
A considerable portion of the drag for a bluff body commercial data sets. However, the correlation of boattail aerodynamics with
vehicle is attributed to low pressures acting on the base area. Reynolds number is not as straightforward as the analysis presented
However the complexity of a commercial vehicle and the local in the early sub-sections of this Chapter. Boattail aerodynamics is
geometric features of the aft most portion of this class of vehicle can influenced by the boundary layer condition and thickness, vehicle
differ considerably resulting in large variations in separation patterns base area, turning angle and wake characteristics. To support this
and base flows making the study of drag trends a challenge. analysis a number of relevant papers were reviewed. Bluff base and
boattail aerodynamic data and information can be found in references
Recognizing the complexity of base drag problem previous studies 17, 25, 26, 27, 28, 45, 51, 55, 153, 154, 159, 160, 161, 162, 163, 170,
have employed a simplified host bluff-body that is streamwise 171, 172, 187, and 211, 212, 213, 214. Data from the subject
continuous. Most notable of these studies is the work of Hoerner in references showed inconsistent trends with Rew, as represented by the
1965 [38] and a modification of Hoerner's work by Butsko in 1965 data shown in figure 3.10. These results are from three subscale fixed
[209]. Hoerner recognized that base drag is inversely proportional to floor wind tunnel tests for Rew values of 1.5, 2.3 and 5.0 million in
the thickness of the boundary layer immediately upstream of the base which the drag reduction and optimum boattail angle vary randomly
where this boundary layer is a function of upstream body length, with Rew. The observed variation is not surprising given that boattail
body surface roughness and surface shape. Hoerner states that the performance requires that the flow transitioning from the upstream
boundary layer serves as an “insulating sheet” between the external portion of the vehicle onto the angled boattail must remain attached.
flow “jet pump effect” and the dead air wake. He further states that The ability of the flow to negotiate the expansion onto the boattail
the boundary layer thickness is proportional to the forebody drag surface is a function of the boundary layer state, thickness of the
(CD,F), which is described as “skin drag”. From this analysis Hoerner boundary layer, boattail angle and the base area size compared to the
developed an empirical relationship that represents the ratio of boundary layer thickness. The relationship between boundary layer
average base pressure (CP,B) to forebody drag, see equation 3.1. thickness and the size of the base area determines the extent that
Equation 3.2 represents the adjustment made by Butsko and equation corner flow will introduce three dimensional flow features that may
3.3 is Hoerner's relationship for two-dimensional bodies. interfere with the optimal boattail flow characteristics.

(3.1)
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 615

Figure 3.11. Effect of boundary layer thickness on Boattail drag reduction.


Figure 3.10. Effect of Reynolds number on optimum Boattail angle.
3.2.4. Wheel and Tire Flows
Based on the above arguments, boundary layer characteristics were
chosen as the correlation parameter for boattail performance. This The variation in wheel/tire CD with Reynolds number was viewed as
approach is supported by the work of Hoerner who suggest a a parameter of interest due to the exposure of commercial vehicle
relationship between vehicle base drag and boundary layer thickness wheel/tire to the free stream flow resulting in potentially large drag
that is a function of transition location. The data analysis focused on values. Unfortunately the relevant data for this component is limited
vehicles with forward body sections that have minimal pressure to isolated wheel/tire studies (without the presence of a vehicle), see
gradient, such as a box trailer. This allowed the use of flat plate references 11, 22, 30, 33, 61, and 214, 215, 216, 217, 218, 219, 220,
analysis of boundary layer transition and boundary layer thickness 221, 222, 223, 224, 225. On a positive note, these data are for
estimates. Presented in figure 3.11 are analysis results showing the large-scale isolated wheel/tire test articles where full-scale Reynolds
variation in the fractional change in CD with increasing non- number sensitivities have been defined. The data show that the
dimensional boundary-layer thickness. The non-dimensional variation in drag with Reynolds number is equivalent for static and
boundary layer thickness parameter (δA) is the ratio of boundary rotating wheels. For the current investigation the original test
layer thickness, immediately upstream of the boattail, to the square Reynolds number values have been converted to Reynolds number
root of the base area. The boundary layer thickness calculation is for values based on the width of a typical commercial vehicle. Presented
a flat plate boundary layer and assumes a boundary layer transition in figure 3.12 is representative data from Cogotti [22] showing the
Reynolds number (Reblt,x) value of 1.0 million. Note; all data variation in CD with Rew for two isolated wheel/tire configurations.
presented are for conditions in which boundary layer transition These data show a Retr,w of approximately 2.0 million, consistent
occurred within the first 30 percent of the test article length. Results with the edge radius data presented previously in sub-section 3.2.3. In
are presented for sub-scale wind tunnel test, full-scale wind tunnel addition the tire data show that aerodynamic studies of commercial
test, and coastdown tests. Figure 3.11 show that boattail performance vehicles with rotating wheels should be performed at Retr,w values
decreases with increasing δA. Noted on the figure are values of δA for greater than 2.0 million in order to replicate operational conditions.
a 53-foot trailer at speeds of 60 and 15 mph, representing full scale
Reynolds number values of 4.6 and 1.2 million, respectively. These
reference conditions indicate boattail effectiveness will reduce by 50
percent during a typical coastdown test or for a subscale wind tunnel
test compared to full scale Reynolds number test. These data
highlight the challenge associated with low Reynolds number base
drag studies that require full scale modeling of both the boundary
layer state and boundary layer thickness. Also shown are reference
points for 42 and 26-foot trailers at 60mph. These data show that the
thinner boundary layer, due to the shorter vehicle length, could
increase the drag reduction benefit of a boattail. To ensure test
results represent operational conditions aerodynamic studies of bluff
base commercial vehicles, including base flows and base mounted
drag reduction devices, shall be performed at Retr,w values greater
than 2.0 million and should include a Reynolds number sweep range
of at least 2.0 million.
Figure 3.12. Variation in CD with Reynolds number for two isolated wheels
geometries [22].
616 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

3.3. Commercial Ground Vehicles with Rew for three yaw angles. The figure shows that increasing Rew
Aerodynamic similarity requires that Reynolds number, flow from 0.5 to 6.4 million results in a 4 percent reduction in CD at 0
conditions and geometric details for the test article match the degree yaw, 7 percent reduction in CD at 5 degree yaw, and 11 percent
operational conditions for the full-scale commercial vehicle [3, 40, reduction in CD at 10 degree yaw. Associated with the drag reduction
143]. In practice, satisfying this requirement can be difficult for wind is an increase in the transcritical Reynolds number value with
tunnel testing, computational studies, as well as coastdown testing. increasing yaw angle. The Retr,w value varies from 3×106 at 0 degrees
Although Reynolds number can be matched in computational work yaw to 5×106 at 10 degrees yaw.
there is typically little information known about the state of the
boundary layer on the full-scale reference vehicle to ensure that all
relevant flow features are accurately simulated in the computational
analysis. Coastdown testing has a different challenge due the large
variation in Reynolds number during a test run. This variation will
produce significant changes in the boundary layer and base separation
characteristics that will impact the CD values resulting and increase
data uncertainty. An overarching challenge in defining the Reynolds
sensitivity for a complete vehicle is accounting for the impact of
these differences on the various components of the vehicle.

Vehicle aerodynamic data and information can be found in references


1, 2, 4, 12, 13, 15, 17, 18, 19, 20, 21, 23, 24, 25, 27, 37, 39, 40, 41,
42, 46, 52, 53, 55, 57, 60, 65, 66, 153, 160, 164, 165, 166, 167, 168,
169, 171, 173, 174, 176, 177, 181, 183, 185, 186, 187, 188, 189, 190,
191, 192, 212, 213, 222, and 226, 227, 228, 229, 230.
Figure 3.14. Variation in CD with Reynolds number for a generic combination
In addition to the variation in Reynolds number sensitivity for each vehicle [2].
vehicle component this class of vehicle has a significant Reynolds
number sensitivity with changes in yaw angle. An example of the In 2013 McArthur [66] obtained drag data for sub-scale models of a
yaw sensitivity is clearly seen in the Storms [55] data shown in figure current model cab-over design and a generic bluff body over a
3.13. While the test article in the Storms study was a generic width-based Reynolds number range of 0.27 to 2.40 million, as
combination vehicle it does have representative aerodynamic shaping shown in figure 3.15. Each test article was 0.33 scale and the models
similar to post 2000 commercial vehicles. The oval shaded regions on were positioned over a raised floor and mounted to an under-floor
the right side of the figure highlight the increasing increment in CD balance. The figure shows that all of cab-over data is in the
due to increasing Rew for yaw angle of 0, 5 and 10 degrees. transitional region and the generic body data extends from the
transitional region to the transcritical region with a transcritical
Reynolds number of 1.2 million. The cab-over model data shows a
nearly linear decrease in CD of 10 percent as Rew is increased from
0.27 to 1.1 million. This drop in drag is followed by an increase in CD
of 5 percent as Rew is increased further to 2.4 million. The cab-over
data never achieves the transcritical Reynolds number, as seen for the
generic body model, indicating the minimum Reynolds number (i.e.
transcritical Reynolds number) is greater than 2.4 million. The
generic bluff body model exhibits a 22 percent decrease in CD as Rew
is increase from 0.27 to 1.2 million at which point it remains constant
for all further increases in Rew. These data indicate the minimum test
Reynolds (i.e. transcritical Reynolds number) for the generic model is
1.2 million. The difference in the minimum Reynolds number is most
likely attributed to the significant difference in geometric complexity
of the two models. For the cab-over design the more complex
variation in drag with Reynolds number is due to the difference in
sensitivity of the tractor and trailer as well as the local effects of the
Figure 3.13. Effect of Reynolds number on Yaw angle dependent CD for a
various model components to changes in Reynolds number. This is in
generic combination vehicle [55].
contrast to the generic model whose variation in drag is dictated by
In 2012 Wood [2] analyzed Storm's data and derived a minimum two primary geometric features, the local curvature of the front
width-based Reynolds number value between 3.0 and 5.0 million, see section and the length of the body.
figure 3.14. Presented in figure 3.14 is the fractional change in ΔCD,
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 617

number sensitivities. Using the argument presented for the data of


figure 3.15 we can estimate that testing any scale of the VNL630® in
the transitional range results in a maximum uncertainty greater than
8.3 percent at an Rew of approximately 1.0 million and greater than
3.3 percent at an Rew of approximately 2.0 million.

Figure 3.15. Variation in drag with width based Reynolds number [66].

Shown in figures 3.16, 3.17, 3.18, 3.19 are wind tunnel test results by
Leuschen [65] for four current-model heavy trucks, for Rew from 1.0
to 6.8 million. Figures 3.16 and 3.17 present curve fits of the
Leuschen data that show the variation in drag coefficient with
increasing Reynolds number for the zero degree yaw condition. The
Figure 3.16. Variation in drag for full-scale vehicles with width based
drag and Reynolds number values presented for each vehicle are
Reynolds number [65].
referenced to the vehicle's value for a test speed of 100km/hr (62
mph). Yaw and surface pressure test results for the full-scale
ProStar® vehicle are shown in Figures 3.18 and 3.19, respectively.

Figure 3.16 compares results for four full-size operational vehicles.


The test Reynolds number range, for all vehicles, includes the
transitional Reynolds number region, the transcritical Reynolds
number value and the transcritical Reynolds number region. The data
indicate that the critical Reynolds number and the subcritical
Reynolds number region occur at Rew below 1.0 million. The data
show that the transcritical Rew for all vehicles is 80 percent of the
100km/hr Rew value, which corresponds to a Rew of approximately
3.8 million. The differences in the vehicle specific variation in
transitional Reynolds number range highlights the complexity of the
flow over this class of vehicle as well as the variation in Reynolds
number sensitivity for each vehicle component. Realizing that the
source of the variation in the transitional range is unknown and that Figure 3.17. Comparison between full- and half-scale VNL630® vehicles of
the data suggest the spread in the curves will increase with decreasing the variation in drag with width based Reynolds number [65].
Reynolds number we can assume that the minimum range of possible
values is the difference between the ProStar® and VNL630® curves Commercial vehicles are known to exhibit significant variations in
at the lowest Reynolds number value tested (i.e approximately 1.0 drag with yaw angle. An example of this effect is presented in figure
million) corresponds to a value of 0.05. Assuming a vehicle CD value 3.18 for the full-scale ProStar® vehicle at four Reynolds numbers.
of 0.60 we can estimate that testing heavy commercial trucks in the Leuschen [65] indicates that the CD values shown have been offset by
transitional range results in a minimum uncertainty greater than 8.3 a fixed value for proprietary concerns however the increments are as
percent at an Rew of approximately 1.0 million and greater than 3.3 measured. At operational conditions (Rew/Rew,FS = 1.0) the data show
percent at an Rew of approximately 2.0 million. that increasing yaw angle to 7.5 degree results in a CD increase of
0.09, however, decreasing Reynolds number produces mixed results.
Scale effects are presented in figure 3.17 for a full-scale vehicle and For 0 degree yaw angle the CD value increases while for yaw angles
half-scale model of the VNL630®. The data show the full-scale and of 5 degree and 7.5 degree the CD value decrease. This trend is
half-scale test articles have opposite trends as a function of Reynolds inconsistent with that seen for the generic vehicle model of Storms
number. Leuschen [65] suggests that the opposing trends may be due [55] in figure 3.13. Despite the differences between the Storms and
to differences in; geometry, underhood cooling system modeling or Leuschen data the general conclusion that Reynolds number
model support hardware. Regardless of the root cause, this result sensitivities are significantly affected by yaw angle is consistent.
reinforces the importance of understanding not only the source but
also the overarching fluid dynamics associated with the Reynolds
618 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

Considering the data shown in figures 3.13, 3.14, 3.15, 3.16, 3.17,
3.18, 3.19, a conservative estimate of the minimum Reynolds number
(i.e. transcritical Reynolds number) to model the aerodynamics of a
full-scale commercial vehicle is 3.0 million. The data also show that
the Retr,w for wind tunnel shall be determined based on a detailed
analysis over the yaw angle for the investigation. Similarly, for
coastdown testing the value of Retr,w shall be based on a yaw
dependent minimum Retr,w assessment. See Chapters 4 and 5 for
additional guidance.

3.4. Summary Findings


Bluff body aerodynamics has been reviewed and discussed in
increasing levels of complexity from simple two- and three-
dimensional shapes, to major vehicle components and finally related
to sub-scale and full-scale vehicles. Wind tunnel data served as the
Figure 3.18. Variation in drag of the ProStar® vehicle with yaw angle, for source for all discussions. The terminology of Roshko was introduced
several width based Reynolds numbers [65]. and adopted for discussing that the variation in drag with Reynolds
number in which the subcritical, critical, transitional and transcritical
A final set of Leuschen's [65] data to discuss is surface pressure
flow regimes were defined. Also discussed was the impact of
measurements for the ProStar® full-scale vehicle presented at 0
boundary layer state on the aerodynamics of this vehicle class.
degree yaw angle, see figure 3.19. The data were obtained along the
vehicle upper surface centerline including the tractor base, trailer
Data for simple two- and three-dimensional shapes show that the
front and trailer base. Pressure data were also obtained along each
transcritical Reynolds number can vary over a broad Reynolds
side of the tractor cab. The specific location of each pressure
number range for small changes to a vehicle component. In contrast
measurement is noted in the lower portion of the plot. Measurements
vehicle component data as represented by edge radius, boattail and
are shown for Rew/Rew,FS values of 0.8 and 0.4 which are for Rew of
wheels all showed a transcritical Reynolds number greater than 2.0
approximately 3.8 and approximately 1.9 million, respectively. The
million. A summary of the vehicle data identifies the minimum
pressure data is presented as the difference between the value at
transcritical Reynolds number as 3.0 million. The vehicle data also
Rew,FS and that measured at a lower Rew. The data show minor
highlighted the complexity and Reynolds number dependence of the
difference along the length of the vehicle for Rew/Rew,FS = 0.8 but
flow over a commercial vehicle and clarified the importance of
significant differences for Rew/Rew,FS = 0.4. Compared to the higher
modeling the full-scale aerodynamics as well as on-body and
Reynolds number data the results for Rew/Rew,FS = 0.4 show the
off-body flow conditions when conducting experimental and
pressure on the front of the tractor and trailer is lowered which would
computational studies at sub-scale or off-design conditions.3.5.
produce a drag reduction however the pressure on the base of the
tractor and trailer is also lowered which would result in a drag
increase. Unfortunately the sparse data set does not allow for a 3.5. References
adequate analysis that supports or disputes the increase in drag with a 156. Ahmadi, M. and Garry, K., “Preliminary Investigation of the Influence
decrease in Reynolds number shown in figure 3.16. However the data of a Ground-Plane Boundary Layer on the Aerodynamic Characteristics
of Road Vehicle Models Tested Over a Fixed Ground,” SAE Technical
does support the need for detailed understanding of Reynolds number Paper 960675, 1996, doi:10.4271/960675.
sensitivities for this class of vehicle. 157. Ahmed, S., Ramm, G., and Faltin, G., “Some Salient Features Of The
Time-Averaged Ground Vehicle Wake,” SAE Technical Paper 840300,
1984, doi:10.4271/840300.
158. Alam, F., Watkins, S., Zimmer, G., and Humphris, C., “Effects of Vehicle
A-pillar Shape on Local Mean and Time-Varying Flow Properties,” SAE
Technical Paper 2001-01-1086, 2001, doi:10.4271/2001-01-1086.
159. Al-Garni, A., Bernal, L., and Khalighi, B., “Experimental Investigation
of the Near Wake of a Pick-up Truck,” SAE Technical Paper 2003-01-
0651, 2003, doi:10.4271/2003-01-0651.
160. Bayraktar, I., Landman, D., Cary, W., Wood, R. et al., “An Assessment
of Drag Reduction Devices for Heavy Trucks Using Design of
Experiments and Computational Fluid Dynamics,” SAE Technical Paper
2005-01-3526, 2005, doi:10.4271/2005-01-3526.
161. Lienhart, H. and Becker, S., “Flow and Turbulence Structure in the Wake
of a Simplified Car Model,” SAE Technical Paper 2003-01-0656, 2003,
doi:10.4271/2003-01-0656.
162. Buresti, G., Fedeli, R., and Ferraresi, A., “Influence of Afterbody
Rounding on the Pressure Drag of an Axisymmetric Bluff Body,”
Journal of Wind Engr. and Ind. Aerodynamics,” 69-71; 179-188. 1997.
163. Buresti, G., “Bluff-Body Aerodynamics,” International Advanced School
on Wind-Excited and Aeroelastic Vibration of Structures,” Lecture Notes,
Figure 3.19. Variation in surface pressure of the ProStar® vehicle with width Department of Aerospace Engineering, University of Pisa, Italy, 2000.
based Reynolds number [65].
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 619

164. Burgin, K., Adey, P. and Beatham, J., “Wind Tunnel Tests on Road 188. Wacker, T., “A Preliminary Study of Configuration Effects on the Drag
Vehicle Models Using a Moving Belt Simulation of Ground Effect,” of a Tractor-Trailer Combination,” Thesis- The University of British
Journal of Wind Engr. and Ind. Aerodynamics, 22 (2-3); 227-236, 1986. Columbia, Department of Mechanical Engineering, October 1985.
165. Cooper, K., Mason, W., and Bettes, W., “Correlation Experience with the 189. Watkins, S., Hoffman, P. and Saunders, J., “Comparison of On-Road
SAE Wind Tunnel Test Procedure for Trucks and Buses,” SAE Technical and Wind-Tunnel Tests for Rigid Truck Aerodynamic Devices,” 9th
Paper 820375, 1982, doi:10.4271/820375. Australasian Fluid Mechanics Conference, Auckland, December 8-12,
166. Cooper, K., “The Wind Tunnel Simulation of Surface Vehicles,” Journal 1986.
of Wind Engineering and Industrial Aerodynamics 17(2):167-198, 1984 190. Watkins, S., “Wind Tunnel Modeling of Vehicle Aerodynamics: With
167. Cooper, K., “The Wind Tunnel Simulation of Wind Turbulence for Emphasis on Turbulent Wind Effects in Commercial Vehicle Drag,”
Surface Vehicle Testing,” Journal of Wind Engineering and Industrial PhD Thesis, Victorian University of Technology, Department of
Aerodynamics 38(1): 71-81, 1991. Manufacturing and Process Engineering, Nov 1990.
168. Cooper, K., “Bluff-Body Aerodynamics as Applied to Vehicles,” Journal 191. Watkins, S. and Cooper, K., “The Unsteady Wind Environment
of Wind Engineering and Industrial Aerodynamics 49(1-3): 1-21, 1993. of Road Vehicles, Part Two: Effects on Vehicle Development and
Simulation of Turbulence,” SAE Technical Paper 2007-01-1237, 2007,
169. Cooper, K., “Truck Aerodynamics Reborn - Lessons from the Past,”
doi:10.4271/2007-01-1237.
SAE Technical Paper 2003-01-3376, 2003, doi:10.4271/2003-01-3376.
192. Wong, H., Cox, R. and Rajan, A., “Drag Reduction of Trailer-Tractor
170. El-Alti, M., Chernoray, V., Jahanmiri, M. and Davidson, L.,
Configuration by Aerodynamic Means,” Journal of Wind Eng. And Ind.
“Experimental and Computational Studies of Active Flow Control on a
Aerodynamics 9(1-2):101-111, Nov. 1981.
Model Truck-Trailer,” 2011 Proc. of the Int.
193. Kettinger, J., “Tractor-Trailer Fuel Savings with an Aerodynamic Device
171. Englar, R., “Advanced Aerodynamic Devices to Improve the
-A Comparison of Wind Tunnel and On-Road Tests,” SAE Technical
Performance, Economics, Handling and Safety of Heavy Vehicles,” SAE
Paper 820376, 1982, doi:10.4271/820376.
2001-01-2072. Government/Industry Meeting, Washington DC, May
14-16, 2001. 194. Van Raemdonck, G. “Design of Low Drag Bluff Vehicles,” PhD Thesis,
University of Delft, Department of Flight Performance and Propulsion,
172. Funderburk, R. and Carper, H., “Wind Tunnel Investigation of an
2012.
Inflatable Aerodynamic Boattail for Tractor-Trailers,” SAE Technical
Paper 960908, 1996, doi:10.4271/960908. 195. Littlewood, R. “Novel Methods Of Drag Reduction For Squareback
Road Vehicles,” PhD Thesis, Loughborough University, Aeronautical
173. Hammache, M., Michaelian, M., and Browand, F., “Aerodynamic Forces
and Automotive Engineering, 2013.
on Truck Models, Including Two Trucks in Tandem,” SAE Technical
Paper 2002-01-0530, 2002, doi:10.4271/2002-01-0530. 196. Hjelm, L. and Bergqvist, B. “European Truck Aerodynamics-A
Comparison Between Conventional and CoE Truck Aerodynamics and a
174. Kirsch, J., Garg, S., and Bettes, W., “Drag Reduction of Bluff
Look into Future Trends and Possibilities,” The Aerodynamics of Heavy
Vehicles With Airvanes,” SAE Technical Paper 730686, 1973,
Vehicles II: Trucks, Buses, and Trains, Lecture Notes in Applied and
doi:10.4271/730686.
Computational Mechanics 41:469-477, 2009.
175. Krajnovic, S. and Davidson, L., “Numerical Study of the Flow Around
197. Newnham, P., “The iInfluence of Turbulence on the Aerodynamic
a Bus-Shaped Body,” ASME Journal of Fluids Engineering 125(3):500-
509, 2003. Optimization of Bluff Body Road Vehicles”. Thesis May 2007.
Loughborough University
176. Landman, D., Cragun, M., McCormick, M., and Wood, R., “Drag
198. Heft, A., Indinger, T., and Adams, N., “Introduction of a New Realistic
Reduction of a Modern Straight Truck,” SAE Int. J. Commer. Veh.
4(1):256-262, 2011, doi:10.4271/2011-01-2283. Generic Car Model for Aerodynamic Investigations,” SAE Technical
Paper 2012-01-0168, 2012, doi:10.4271/2012-01-0168.
177. Landman, D., Wood, R., Seay, W., and Bledsoe, J., “Understanding
199. Storms, B. “A Summary of the Experimental Results for a Generic
Practical Limits to Heavy Truck Drag Reduction,” SAE Int. J. Commer.
Veh. 2(2):183-190, 2010, doi:10.4271/2009-01-2890. Tractor-Trailer in the Ames Research Center 7- by 10-Foot and 12-Foot
Wind Tunnels”. NASA TM 2006-213489, 2006.
178. Lehmukhl, O., Rodriguez, I., Baez, A., Oliva, A. and Perez-Segarra,
200. Platt, R. “Turbulence Factors of N.A.C.A. Wind Tunnels as Determined
C., “On the large-Eddy Simulations for the Flow Around Aerodynamic
Profiles Using Unstructured Grids,” Computers & Fluids 84:176-189, Sphere Tests,” NACA TR 558, 1936.
2013. 201. Cooper, K. R. and Campbell, W. F., “An Examination of the Effects of
Wind Turbulence on the Aerodynamic Drag of Vehicles,” Journal of
179. Lesieur, M. and Metais, O., “New Trends in Large-Eddy Simulations of
Wind Engineering and Industrial Aerodynamics, 9(1-2):167-180, 1981.
Turbulence,” Annual Review of Fluid Mechanics, 28:45-82, 1996.
202. Piomelli, U.: “Large-Eddy Simulation: Present State and Future
180. Newnham, P., Passmore, M., Howell, J., and Baxendale, A., “On the
Directions,” AIAA Paper 98-0534, 36th AIAA Aerospace Sciences
Optimisation of Road Vehicle Leading Edge Radius in Varying Levels
Meeting and Exhibit, Reno, NV, 1997.
of Freestream Turbulence,” SAE Technical Paper 2006-01-1029, 2006,
doi:10.4271/2006-01-1029. 203. Engineering Sciences Data Unit, Wind Engineering, Vol. 26. “Fluid
Forces and Moments on Rectangular Blocks”, ESDU data item 71016,
181. Ortega, J., Salari, K., Brown, A., and Schoon, R.; “Aerodynamic Drag
ISBN: 978-1-86246-372-1, 1971.
Reduction of Class 8 Heavy Vehicles: A Full-Scale Wind Tunnel Study,”
DOE LLNL-TR-628153, 2013. 204. Saltzman, E., Myer, R. Jr., and Lux, D., “Drag Reductions Obtained by
Modifying a Box-Shaped Ground Vehicle,” NASA TM-X-56027, 1974.
182. Polhamus, E., “Effect of Flow Incidence and Reynolds Number on Low-
Speed Aerodynamic Characteristics of Several Noncircular Cylinders 205. Muirhead, V., “An Investigation of Drag Reduction on Box-Shaped
with Applications to Directional Stability ad Spinning,” NACA TN Ground Vehicles,” NASA CR-148829, 1976
4176. Jan. 1958. 206. Saltzman, E., “A Summary of NASA Dryden's Truck Aerodynamic
183. Roy, C., Payne, J. and McWherter-Payne, M., “RANS Simulations Reasarch,” SAE Technical Paper 821284, 1982, doi:10.4271/821284.
of a Simplified Tractor/Trailer Geometry,” ASME Journal of Fluids 207. Saltzman, E. and Meyer, R., “A Reassessment of Heavy-Duty Truck
Engineering 128(5):1083-1089, 2006. Aerodynamic Design Features and Priorities,” NASA TP-1999-206574.
184. Sakuma, Y. and Ido, A., “Wind Tunnel Experiments on Reducing 1999
Separated Flow Region Around Front Ends of Vehicles on Meter-Gauge 208. Diebler, C. and Smith, M. “A Ground-Based Research Vehicle for Base
Railway Lines,” Railway Technical Research Institute, Quarterly Report Studies at Subsonic Speeds,” NASA TM 2002-210737, 2002
50(1):20-25, 2009. 209. Butsko, J., Carter W., and Herman W., “Development of Subsonic Base
185. Smith, G., “Commercial Vehicle Performance and Fuel Economy,” SAE Pressure Prediction Methods,” AFFDL-TR-65-157, Vol. 1, Aug. 1965.
Technical Paper 700194, 1970, doi:10.4271/700194. 210. Gillieron, P. and Spohn, A., “Flow Separations Generated by a
186. Steers, L., Montoya, L., and Saltzman, E., “Aerodynamic Drag Simplified Geometry of an Automotive Vehicle,” IUTAM Symp.,
Reduction Tests on a Full-Scale Tractor-Trailer Combination and a Unsteady Separated Flows, 2002.
Representative Box-Shaped Ground Vehicle,” SAE Technical Paper 211. Schembri-Puglisevich, L., “Large Eddy Simulation for Automotive
750703, 1975, doi:10.4271/750703. Vortical Flows in Ground Effect,” PhD Thesis, Loughborough
187. Storms, B., Ross, J., Heineck, J., Walker, S., Driver, D. and Zilliac, G., University, Aeronautical and Automotive Engineering, 2013.
“An Experimental Study of the Ground Transportation System (GTS) 212. Storms, B., Satran, D., Heineck, J., and Walker, S., “Detailed
Model in the NASA Ames 7-by 10-Ft Wind Tunnel,” NASA/TM-2001- Experimental Results of Drag-Reduction Concepts on a Generic Tractor-
209621, February 2001 Trailer,” SAE Technical Paper 2005-01-3525, 2005, doi:10.4271/2005-
01-3525.
620 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

213. Storms, B. and Ross, J., “Aerodynamic Drag Reduction of the 222. Skea, A., Bullen, P., and Qiao, J., “CFD Simulations and Experimental
Underbody of a Class-8 Tractor-Trailer,” SAE Technical Paper 2006-01- Measurements of the Flow Over a Rotating Wheel in a Wheel Arch,”
3532, 2006, doi:10.4271/2006-01-3532. SAE Technical Paper 2000-01-0487, 2000, doi:10.4271/2000-01-0487.
214. Axon, L., “The Aerodynamic Characteristics of Automobile Wheels - 223. Régert, T. and Lajos, T., “The Effect of Wheels and Wheelhouses on
CFD Prediction and Wind Tunnel Experiment,” PhD Thesis, Cranfield the Aerodynamic Forces Acting on Passenger Cars,” Proceedings of
University, College of Aeronautics, 1999. MICROCAD Conference pp. 61-66, Miskolc, Hungary, 2006.
215. Chen, H, Teixiera, C. and Molvig, K., “Realization of Fluid Boundary 224. Wäschle, A., Cyr, S., Kuthada, T., and Wiedemann, J., “Flow around an
Conditions via Discrete Boltzmann Dynamics,” International Journal of Isolated Wheel - Experimental and Numerical Comparison of Two CFD
Modern Physics C 9(8):1281-1292, 7th Int. Conf. on Discrete Simulation Codes,” SAE Technical Paper 2004-01-0445, 2004, doi:10.4271/2004-
of Fluids, University of Oxford, 1998. 01-0445.
216. Hinson, M., “Measurement of the Lift Produced by an Isolated, Rotating 225. Whitbread, L., “Measurement of the Lift Distribution on a Rotating
Formula One Wheel Using a New Pressure Measurement System,” Wheel,” Master's Thesis, Cranfield University, 2000.
Master's Thesis, Cranfield University, 1999. 226. Broughton C., Rainbird W. and Kind R., “An Experimental Investigation
217. Knowles, R., “Monoposto Racecar Wheel Aerodynamics: Investigation of Interference Effects for High Blockage Bluff Bodies in a Slotted-
of Near Wake Structure & Support-Sting Interference,” PhD Thesis, Wall Wind Tunnel Test Section;” Journal of Wind Eng. and Ind.
Cranfield University, College of Defense Technology, Department of Aerodynamics 56(1):23-39, 1995.
Aerospace, Power & Sensors, 2005. 227. Ferziger, J. and Peric, M., “Computational Methods for Fluid
218. McManus, J and Zhang, X., “A Computational Study of the Flow Dynamics.” Springer 3rd edition. Springer Science & Business Media,
Around an Isolated Wheel in Contact with the Ground,” ASME Journal ISBN: 978-3-540-42074-3, 2002.
of Fluids Engineering 128(3) : 520-530, 2006. 228. Mihelic, R. and Ellis, M., “Ramifications of Test Track Curves On
219. Mears, A., Dominy, R., and Sims-Williams, D., “The Air Flow About Aerodynamic Prediction for Tractor Trailer Vehicles,” SAE Technical
an Exposed Racing Wheel,” SAE Technical Paper 2002-01-3290, 2002, Paper 2013-01-2460, 2013, doi:10.4271/2013-01-2460.
doi:10.4271/2002-01-3290. 229. Sulitka, M. and Nožička, J., “Aerodynamic Devices to Reduce the Base-
220. Mears, A., Crossland, S., and Dominy, R., “An Investigation into the and Underbody Drag of Semitrailer Unit Czech Technical University in
Flow-Field About an Exposed Racing Wheel,” SAE Technical Paper Prague, AED 2004.
2004-01-0446, 2004, doi:10.4271/2004-01-0446. 230. Sumantran, V. and Sovran, G., “Vehicle Aerodynamics,” SAE PT-49,
221. Rymiszewski, A., “Improving Wheeled Vehicle Water Speed by Means 1996.
of Wheel Shrouding,” US Army Report No. 8088, AD608552, 1963.
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 621

CHAPTER 4. COMPUTATIONAL The issue may be compunded by the availability of user-friendly


SIMULATIONS computational simulation software that can result in complacent
users. All aerodyamic tools, including computational simulations,
The emergence of CFD in commercial vehicle aerodynamic studies
have risks and care should be taken to corroborate results through
began in the 1990s. Since 2010 more than seventy percent of all SAE
alternative means.
commercial vehicle papers and journal articles contained CFD
content and in 2015 ninety percent of the commercial vehicle papers
showcase CFD results. This growth has been fueled by the reduced 4.1. Flow Similarity
cost, shorter solution timeline, increased accessibility and portability “Reynolds number for the numerical model should match the
of results for CFD, compared to wind tunnel testing. The influence of physical model Reynolds number to ensure same environmental
CFD on the commercial vehicle industry extends well beyond conditions.” SAE J2966 [133]
external aerodynamics and includes internal flows, combustion,
hydraulics, and thermal management. In each of these areas Reynolds Setting Reynolds number in the computational simulation does not,
number plays a key role in defining the fluid dynamics and in guiding by itself, result in a matching of Reynolds number. Reynolds number
method application. matching requires flow similarity that requires geometric and
kinematic similarity, both of which may be difficult and/or costly to
Despite the growing presence of CFD there remains a lack of industry achieve for a commercial ground vehicle. Apart from the global
controls and guidelines in place to measure the quality of a Reynolds number, the boundary layer state depends on geometric
computational simulation. This concern has been formally expressed details, which are often difficult to model for a commercial ground
by most of the leading engineering and scientific standards and vehicle. The boundary layer, in turn, takes influence on the flow field.
professional organizations as well as U.S. governmental agencies. To Hence, simplified models can lead to wrong results, even at correct
address this matter the U.S. Defense Modeling and Simulation Office Reynolds numbers.
(DMSO) [231] initiated a program to investigate this topic area in the
late 1980s and in 1998 the AIAA published the first standard for CFD Relative to aerodynamic investigations of air vehicles it requires
use [232] and in 2009 the American Society of Mechanical Engineers greater resources for a computational simulation of a commercial
(ASME) [233] published standards for CFD and Heat Transfer. The ground vehicle to fully satisfy the geometric similarity requirement
limitations of and the primary sources of uncertainty and error in necessary to achieve flow similarity. This is especially true in the
CFD are outlined by Ferziger [227] below: study or Reynolds number sensitivity effects. The problem lies in the
geometric complexity of a commercial ground vehicle compared to
“Numerical solutions of fluid flow and heat transfer problems are that of an aircraft. The outer mold line (OML) of an aircraft is
only approximate solutions. In addition to the errors that might be typically a continuous three-dimensional surface without breaks and
introduced in the course of the development of the solution algorithm, having a very low surface roughness. In contrast a ground vehicle is a
in programming or setting up the boundary conditions, numerical combination of multiple complex three-dimensional components
solutions always include three kinds of systematic errors: where each component is composed of multiple surfaces separated by
distinct discontinuities. These discontinuities may be gaps or steps
• Modeling errors, which are defined as the difference between with dimensions that are on the order of the local boundary layer
the actual flow and the exact solution of the mathematical thickness or greater [234, 235]. On the forward sections of a vehicle
model; the boundary layer is relatively thin increasing the significance of
• Discretization errors, defined as the difference between the exact small geometric details. Flow through gaps and other restricted
solution of the conservation equations and the exact solution of openings are Reynolds number dependent. The effect on the flow
the algebraic system of equations obtained by discretizing these passing over the discontinuities will be affected by; the local surface
equations, and pressure, flow in or out of the gap and characteristics of the boundary
• Iteration errors, defined as the difference between the iterative layer. However, the typical approach used to model a commercial
and exact solutions of the algebraic equations systems.” ground vehicle does not account for these geometric details and flow
details. The lack of geometric similarity will alter the pressure
In 2011 SAE took a first step in addressing CFD validation and distribution, local flow directions and may alter time scale of the flow
verification when they launched a task force that produced thereby impacting the kinematic and dynamic similarity resulting in a
recommended practice J2966 “Guidelines for Aerodynamic local Reynolds number mismatch and possibly a global Reynolds
Assessment of Medium and Heavy Commercial Ground Vehicles number mismatch. It is recognized that a local Reynolds number
Using Computational Fluid Dynamics” in 2013 [133]. The SAE mismatch may be acceptable if it does not affect the investigation of a
J2966 document is structured for use by both a novice and expert and component where the flow about the component has flow similarity.
therefore provides overarching requirements for the aerodynamic
simulation of commerical vehicles. Reynolds number effects, While CFD can simulate the vehicle/model in the wind tunnel or the
boundary layer modeling options and flow separation criteria are not vehicle in operation the principles of flow similarity must be satisfied.
addressed in detail and are left to the interpretation of the user which Computational simulations shall match the local and global Reynolds
can lead to improper modeling and misleading results. At this time number and boundary layer conditions for the aerodynamic model
the J2966 document has yet to gain a foothold in the industry as and flow condition being evaluated. When comparing simulation
reflected in the limited number of citations within SAE publications. results with experimental data or when comparing across
622 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

computational methods the vehicle geometry, test environment and modeling of the governing physics for each flow regime present on
boundary layer flows of the test article should be accurately each vehicle component requires careful selection and application of
represented by the computational model. The computational the computational approach [227].
requirements for modeling the boundary layer and Reynolds number
sensitivities will vary depending on the model/vehicle geometry and
the analysis conditions and requirements. The numerical simulation
4.3. Computational Fluid Dynamic Methods
of the aerodynamic performance of commercial vehicles should be Computational fluid dynamics is an ever-expanding field of study
based on the guidelines found in J2966 [133] and in accordance with requiring knowledge of fluid dynamics, computational science, and
criteria described in references [232, 233, 236]. computer science. Obtaining a relevant solution requires a basic
understanding of fluid dynamics and computational sciences. To
obtain the best available solution requires an ability to make expert
4.2. Flow Regimes judgments in fluid dynamics and computational sciences. This level
For simple shapes or for the study of localized effects a single of expertise is beyond that of the author, as such the following
Reynolds number may serve as the universal non-dimensional discussion relies heavily on publications by Lesieur [177], Ferzinger
parameter that fully describes the type of flow to be modeled. For this [227], Spalart [239, 240, 246], Launder [247], Wilcox [242], Nichols
problem type at low Reynolds number the boundary layer flow is [244], Chen [248], Derksen [249], and Lockard [250].
dominated by the viscous terms and may be considered to be laminar
and the Reynolds number range is defined as subcritical, see Chapter Computational simulations of commercial ground vehicles may be
3 [3, 120, 227]. At high Reynolds numbers, the flow is dominated by obtained with methods based on either the Navier-Stokes (NS)
the inertial term, and the boundary layer is considered to be turbulent equations [227] or the lattice Boltzman equation (LBE) [248]. Both
in nature, and structures are generally referred to as chaotic [141, methods rely on similar turbulence modeling but have unique
237, 238, 239]. High Reynolds number conditions are defined as the advantages and disadvantages when solving the commercial ground
transcritical range, see Chapter 3. In between the two regions the flow vehicle problem. Navier-Stokes based methods are better suited to
is not fully laminar, nor turbulent and is considered the transition steady flows but can handle unsteady flows while LBE is not well
zone from laminar to turbulent and the Reynolds number range is suited for steady flows but is more robust for unsteady flows. The
defined as transitional, see Chapter 3. LBE method is limited to low Mach numbers (significantly below the
incompressible flow upper limit) whereas the NS methods are not
However, the flow regime for a commercial vehicle is not singular Mach number limited. With these differences in mind the remaining
but a combination of multiple regimes where each regime is discussion will characterize the various methods based on boundary
associated with a component or area of the vehicle. It would not be layer modeling, specifically turbulence modeling.
correct to use the vehicle based Reynolds number value to
characterize the flow regime over a vehicle component or the Turbulence is fluid motion that is unsteady, irregular, three
complete vehicle. An understanding of the various flow regimes dimensional, rotational, with rapid mixing and the conversion of
present on a vehicle is required to properly structure the numerical energy into heat due to the viscous stresses. Turbulent motion
simulation. For complex problems there may be significant contains a wide range of eddies in which the largest eddies, with a
differences between local areas of interest and the global flow size on the order of the flow, transport most of the momentum. The
conditions. Flow characteristics may include; two or three smallest eddies result from the viscous forces and produce high
dimensionality, compressible or incompressible, steady or unsteady, frequency fluctuations. Strategies for modeling turbulence vary
separated or un-separated, and laminar or turbulent [240, 241, 242, greatly from the direct simulation of all length and time scales (i.e.
243, 244]. Each of these features is related to the Reynolds number Direct Numerical Simulation) to a modeling of all length and time
and each will characterize the local and global Reynolds number scales with a single average value (i.e. Reynolds Averaged Navier
sensitivities. For commercial ground vehicles operating at highway Stokes). The Reynolds Averaged Navier-Stokes (RANS) [251] are
speeds we typically characterize the dominant global flow regime as derived by separating the turbulent flow into two parts, a mean
incompressible, unsteady, turbulent, and separated but may have local (ensemble) component and a fluctuating component (figure 4.1)
regions of compressible, steady, un-separated, and laminar as well as through a technique known as Reynolds Decomposition. Through this
transitional flow [120, 177, 228]. However, commercial vehicles vary technique an average of the governing flow equations can be derived.
greatly in size and shape and operate over a broad range of Reynolds These equations result in all the effects of the turbulent fluid motion
number resulting in a diverse range of boundary layer, on-body and being lumped into a single term known as the Reynolds stress tensor.
off-body flow regimes. For example, a vehicle may have areas of This term can then be solved directly or with varying degrees of
laminar and transitional boundary layer flow on the forward portion simplification using a wide range of turbulence models. URANS
of the vehicle and forward section of the trailer as well as regions of [252] is RANS with the unsteady terms from the governing equations
unsteady flow, separated flow, vortex flow, shear flow and wake flow. retained to provide a transient solution. Note, turbulence is not
In addition various smaller scale vehicle components such as cab resolved explicitly in RANS or URANS, but employs turbulence
visor, mirrors, and bumper may experience low Reynolds number models. The choice of turbulence model and the near wall modeling
flows. Each of these flow types will vary significantly with variations approach will impact on the representation of boundary layer state
in environmental factors and vehicle operational conditions. Proper and the resulting flow field.
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 623

suggested if the boundary layer is transitional and the flow is


unsteady with significant flow separation. At relevant Reynolds
numbers for commercial vehicles, the linear size of the smallest
turbulent eddies may be only a few microns making DNS impractical.
However, if the problem is related to a vehicle component with
complex flow conditions, and a high degree of confidence is required,
a DNS solution may be warranted.

The use of hybrid methods has grown with many variants available,
Figure 4.1. Time averaging for a statistically steady flow (left) and ensemble including; LES with near-wall modeling, very large eddy simulation
averaging for an unsteady flow (right) (Ferziger [227], Fig 9.10). (VLES) and detached-eddy simulation (DES). In LES, the grid
captures the turbulent kinetic energy and turbulence is modeled in the
Modeling schemes that fall between Direct Numerical Simulation
near wall region [237]. VLES is typically performed on coarser grids
(DNS) [227] and Reynolds Averaged Navier-Stokes (RANS) are
than LES, and as such is computationally less expensive but is more
unsteady RANS (URANS), Large-Eddy Simulation (LES) [252],
dependent on the modeling. DES is a modification of the RANS
Very-Large-Eddy Simulation (VLES) [227], and Detached-Eddy
approach, where the model represents the boundary layer with RANS
Simulation (DES) [252]. Hybrid methods evolved to address the
and large-scale separations with LES. Typically, Navier-Stokes based
inability of turbulence model approximations in RANS to resolve
CFD codes will utilize LES and DES as hybrid methods and lattice
large-scale unsteadiness. In these methods the large-scale turbulent
Boltzmann codes will utilize VLES.
structures are resolved directly through the grid while smaller scales
are modeled. This allows the small-scale turbulence near the body
surface to be modeled while capturing the large-scale unsteady
separation and vortex shedding generated by a bluff body. Shown in
figure 4.2 is a graph reflecting the range of turbulent structures
simulated by the various method types, where the wave number is
inversely proportional to the turbulent length scale. The graph
indicates that RANS models all turbulent structures with a single
length scale. At the other extreme is DNS, which simulates all scales
without any modeling. In between are Hybrid and LES methods
which model the smaller scales by representing the sub-grid
turbulence with a turbulence model to simulate the larger scales with
the Navier-Stokes equations. Not shown but equally important is the
computational cost associated with each simulation type. RANS
would typically be the least expensive with DNS at the opposite end
of the scale with hybrid methods falling between the two.

Figure 4.3. Vorticity isosurfaces by a circular cylinder: ReD = 50,000,


Figure 4.2. Graph depicting the range of turbulent energy spectrum modeled laminar separation. Experimental drag coefficients: a. 0.78, b. 1.73, c. 1.24,
by various turbulent modeling approaches. Nichols [244] d. 1.16, e. 1.26, f. 1.28 [246, 253]

While each method can provide a simulation of the external or A graphical depiction of the difference in flow simulation results
internal aerodynamics of a vehicle it is incumbent on the user to between RANS and methods that simulate the larger turbulent
select the simulation approach best suited for the task at hand [133]. structures is presented in figure 4.3. The six images of figure 4.3
To obtain general force trends a RANS method may be selected for depict RANS, URANS and DES simulations for a circular cylinder
most complex geometries. For a detailed study of a complete vehicle with laminar crossflow. These images are from Spalart’s 2009 paper
a RANS approach may be used if the flow is attached, steady and has [246] in which the images are attributed to Travin [253]. The figure
a turbulent boundary layer. In contrast a hybrid method would be
624 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

shows the inability of RANS and URANS to capture the large-scale In general, for low Reynolds number flow the boundary layers are
turbulent structures but URANS (see images b and c) does capture the thicker and it is possible to use a large number of grid points to
shedding frequency simulated by DES (see images d, e, and f). DES resolve the viscous and transition layer. The boundary layer thickness
simulations show greater sensitivity to grid refinement (see images d is a property of the flow field and a function of Reynolds number. For
and e) and less for sub-grid model choice (see images e and f). higher Reynolds numbers, turbulent boundary layers will be present.
A turbulent boundary layer usually has a very thin viscous sublayer at
When accurate prediction of separated wake flows and the impact on the wall, which behaves laminar, because the flow cannot move
aerodynamic drag are needed, CFD methods which capture the through the wall and hence no three dimensional vortices can be
turbulent, transient flow structures in separated regions using the present in this region. To determine the location of a grid point in the
highest possible fidelity should be used. Methods such as VLES or flow field, a non-dimensional wall distance parameter (y+) can be
DES are recommended over RANS so that turbulent flow structures used, see Eq. 4.1.
are simulated rather than modeled. Specific modeling criteria for
commercial vehicles are found in SAE J2966 [133]
(4.1)

4.4. Boundary Layer Modeling Where:


As discussed in Chapter 2, a boundary layer starts off as laminar and
transitions downstream to fully turbulent flow. The streamwise y is the distance to wall
location of transition onset is a function of Reynolds number, vehicle Uτ is the friction velocity, Uτ = τw0.5ρ−0.5
geometry and pressure gradient. There are situations where all three
states should be represented in the simulation, as described in the ρ is the density
following two narrowly defined cases. Both cases will assume a τw is the wall friction
transition Reynolds number of 1.0 million.
ν is the kinematic viscosity
The first case is a combination vehicle operating at 65 mph with a The turbulent boundary layer velocity profile of a high Reynolds
width based Reynolds number of 5.2 million. Assuming a transitional number flow can be divided into four regions; viscous sublayer,
Reynolds number of 1.0 million tells us that the first 1.6 feet of the buffer layer, log layer and wake, see figure 4.4 [254]. The associated
trailer and tractor boundary layer could be laminar. This length y+ value range for the incompressible velocity profile in each region
corresponds to three percent and ten percent of the trailer and truck of the turbulent boundary layer is noted in figure 4.4. The y+ value
length, respectively. If we discount local Reynolds number effects, where the profile transitions from the inner to the outer layer varies
impact of yaw angle or the effect on the aerodynamics of a specific with the Reynolds number and pressure gradient. The outer layer is
component it is reasonable to assume that the presence of laminar much more sensitive to a pressure gradient.
flow and transition will not greatly impact the vehicle aerodynamics.
If a turbulent boundary layer is to be resolved in detail a y+ value
The second case is for a sub-scale wind tunnel test of the same around 1 is recommended for the first grid point. To reduce
vehicle. Assuming a transitional Reynolds number of 1.0 million and computational cost in Navier Stokes applications a value of y+ < 5 is
the test is performed without forced transition. The model scale is chosen, which is still located within the viscous sublayer [131]. The
0.125 and the test wind speed is 110 mph providing a width based flow in this layer is assumed to be laminar, and viscous stress
Reynolds number of 1.0 million. These conditions produce a laminar dominates the wall shear. As the wall shear is a required input for
length of 1.0 foot that correlates to the first fifteen percent and fifty calculating the absolute size of the first grid cell, setting the right y+
percent of the models trailer and truck lengths, respectively. This case value is an iterative problem and is not known in advance. Therefore,
requires that all boundary layer states be modeled and/or simulated a rough approximation is used as an initial value for the absolute size
and if you take into consideration local Reynolds number effects, of the first cell at the wall (ywall,cell) which is then corrected by the
impact of yaw angle or the effect on the aerodynamics of a specific results of the numerical calculation. In this case;
component the boundary layer simulation requirements increase
significantly. Both cases assume geometric similarity of both the y+ ≈ 1,
vehicle/model and the test environment are satisfied.
when 20/Reu ≤ ywall-cell ≤ 50/Reu
4.4.1. Modeling Low and High Reynolds Number Flows and unit Reynolds number (Reu) = U/ν.
To achieve flow similarity and accurate Reynolds number modeling
the boundary layer must be simulated or modeled accurately. This is To verify whether the boundary layer is resolved correctly requires
especially true for drag prediction and capturing separated flows. an examination of the turbulence transition region, where the
Depending on the simulation objective and the aerodynamic thinnest boundary layer of turbulent type occurs. Here the size of the
phenomena, it is important to understand and select the proper viscous sublayer is the thinnest. The skin friction coefficient rises to
boundary layer modeling/simulation. The first step is to determine if a local maximum and so does y+ for a given wall distance, e.g. y+ for
you have a laminar (low Reynolds number flow) or turbulent the first grid point.
boundary layer (high Reynolds number flow).
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 625

(4.2)

A correlation-based transition model can be used to avoid direct


evaluation of non-local quantities, such as momentum thickness
Reynolds number, Req. However all methods require fine grid to
capture the transition, just not as fine as needed to capture the
turbulent structures seen in the transition.

Until recently the aerodynamics of a commercial vehicle was dictated


by separated flows. As a result the impact of transition modeling on
vehicle aerodynamics had limited value. As the influence of
aerodynamics on ground vehicle shaping and performance grows so
Figure 4.4. Subdivisions in near-wall region of a turbulent boundary layer will the importance of boundary layer flows and transition. At this
plotted in semi-log coordinates (Bird [254] 1960, Figs 5.3-1 and 5.5-3) time there are no published examples of transition modeling on a
commercial vehicle in operation or for a sub-scale test of a
In high Reynolds number turbulent flows, wall boundary layers commercial vehicle model.
become extremely thin with high velocity gradients and a coarse grid
will not produce representative results, see figure 4.5. This problem A demonstration of the importance of transition in predicting
may be avoided by using the “law of the wall” [255] to approximate aerodynamic performance is shown in the study of a multi-element
the boundary layer. This approach uses semi-empirical formulas wing depicted in figures 4.6, 4.7 and 4.8. These results are from
called wall functions, which rely on the existence of a logarithmic Rumsey [262, 263], Eliasson [264] and Shankara [265)] for the 1st
region (i. e. log-law region) in the velocity profile, instead of AIAA High-Lift Prediction Workshop (HLPW-1). The workshop
explicitly modeling this region. These wall functions generally made use of experimental data obtained on the National Aeronautics
assume that the first grid point is within the logarithmic region, and Space Administration (NASA) Trapezoidal wing model [262],
requiring a y+ > 30. This reduces the number of points required to which obtained location of the transition zone, see figure 4.6. Shown
discretize a flow field and reduces solution cost. In regions of in figure 4.7 is the grid used by Shankara [265] RANS analysis of the
separated flows, these wall functions may not be valid and adaptive NASA Trap wing at a Mach number of 0.202 and Reynolds number
wall methods or Low Reynolds wall modeling should be used. based on the mean aerodynamic chord of 4.6 million. A two-equation
predictive transition model was employed in addition to steady-state
RANS. For this study, a y+ between 0.1 and 1.0 was applied, with 30
prismatic layers to model the boundary layer profile.

Figure 4.5. Schematic of turbulent boundary layer depicting relative scale of


viscous sub-layer and buffer layer.
Figure 4.6. NASA Trapezoidal Wing geometry. Rumsey [264].

4.4.2. Modeling Laminar to Turbulent Transition Lift characteristics for the trap wing are shown in figure 4.8, which
Only DNS and LES can capture laminar to turbulence transition in show that modeling transition results in an improved prediction of
the boundary layer explicitly [256, 257, 258, 259]. However both are lift. The use of the transition model enabled prediction of separation
computationally expensive and are not typically used in the study of for steep angle of attacks, which closely matched the experimental
commercial vehicles. The alternative to direct simulation is to make data. The fully turbulent solution (blue line) converged to a maximum
use of a modeling method [257, 260, 261]. When transition is CL value of 2.8, while the transition model (red line) captured the
modeled, its beginning and end point are known. The first step is to locations of separation/transition more accurately and matched the
determine the streamwise position where transition begins, known as experimental lift curve up to 32 degree angle of attack.
transition onset. This is the point where disturbances in the laminar
flow increase such that turbulent spots form indicating the beginning These results reinforce the importance of modeling laminar regions and
of transition. An empirical method to determine transition onset may transition for scale-model wind tunnel testing conducted at reduced
be based on the momentum thickness Reynolds number (Reθ) as Reynolds numbers or when there is a need for precise determination of
defined in Eq. 4.2. forces on vehicle components with laminar flow regions.
626 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

Figure 4.9. LES grid for generic bluff body, Eichinger [266].

Figure 4.7. Section views of the mesh structure. Shankara [265]

Figure 4.8. Lifting characteristics of wind body model. Rumsey [262] and
Shankara [265]

4.5. Modeling Turbulent Wake


The work of Eichinger [266] will be used to highlight the importance
of unsteady separations for commercial vehicles, see figures 4.9, 4.10
and 4.11. The generic bluff body shown in figure 4.9 has been subject
to extensive study and allows for the selection of an effective
empirical modeling of the boundary layer characteristics and
transition location in RANS. The body had a height to width ratio of
0.74 and length to width ratio of 2.68. Eichinger used the
experimental data of Ahmed [157], Becker [161], and Gillieron [210]
for comparison to LES and RANS results. Depicted in Figure 4.9 is
the LES grid with every other grid point removed. The LES grid was
minimized from the optimum to reduce cost but maintained 24.9
million hexahedral cells. The RANS simulations were obtained with
a mesh of 11.9 million cells. LES and RANS simulations were
obtained at a Reynolds number of 500,000 based on body length (Rew
= 187,000). The reduced Reynolds number minimized the
computational cost of the LES solution while not impacting the
unsteady base flow and loading characteristics.
Figure 4.10. LES and RANS predicted surface pressures and streamlines at 0
degree yaw, Eichinger [266].
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 627

LES and RANS results are shown in Figures 4.10, and 4.11. Surface
pressure coefficient contours are shown in Figure 4.10a and 10b for
LES and 4.10c and d for RANS. Overlayed on the RANS surface
pressure contours in Figure 4.10c and d are the time averaged LES
streamlines (white lines). The streamlines show a laminar separation
and reattachment on the front side and top edges of the body with the
RANS transition model producing a more forward laminar separation
and reattachment. LES and RANS predicted drag forces on the
bluff-body front differ significantly with the RANS producing a c. RANS Side View, vertical centerline cut
positive drag coefficient (0.027) and the LES showing a negative drag
coefficient (−0.016). A more pronounced difference between the
RANS and LES results are seen in the base surface pressure
coefficient plots of Figures 4.10b and 10d as well as the wake flow
images of Figure 4.11. Reviewing the results of Figure 4.10 and 4.11
show that the RANS and LES freestream stagnation (Fs) and rear
stagnation (Sr) points differ greatly and the RANS does not predict
the base secondary vortices located at the side and top edge, as seen
in the LES solution. The LES predicted base drag is twenty-six
percent greater than that for RANS.
d. RANS Top View, horizontal centerline cut
Simulations performed on the Generic Conventional Model (GCM)
by Horrigan et al. [267] confirmed that an unsteady CFD model was Figure 4.11. LES (time averaged) and RANS predicted wake streamlines at 0
capable of predicting separated wake regions and the base pressure degree yaw. Eichinger [266]
distribution as measured in the NASA Ames 12 foot pressure wind
tunnel. In this case, a Lattice Boltzmann solver was used with a
VLES turbulence model. The base pressure distribution was
compared for various yaw angles at a Rew value of 6.2 million.
Particle Image Velocimetry (PIV) measurements at a reduced Rew
value of 1.1 million in the tractor trailer gap and wake regions were
also compared to simulation results. The PIV measurement
uncertainty of instantaneous velocity measurements was stated as 2
percent for in-plane components and 4 percent for the cross-plane
component, Heinecke [268]. The comparison between experimental
results and simulations are shown in Figures 4.12 and 4.13.

a. Base pressure distribution on centerline of trailer base

a. LES Side View, vertical centerline cut (time avg)

b. Static pressure coefficient on trailer base

Figure 4.12. CFD simulation results of the base pressure distribution on the
b. LES Top View, horizontal centerline cut (time avg)
GCM model using a Lattice Boltzmann transient solver at Rew of 6.2 million,
compared to experiments of the NASA Ames 12 foot tunnel Heinecke [268].
628 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

transitional boundary layer flow on its forward portion and forward


section of the trailer as well as regions of unsteady flow, separated
flow, vortex flow, shear flow and wake flow. Proper modeling of the
governing physics for each flow regime present on each vehicle
component requires careful selection and application of the
computational approach. These concerns reinforce the importance to
model laminar regions, using either a correlation with scale-model
wind tunnel testing conducted at reduced Reynolds numbers or using
a precise measurement of forces on certain vehicle components with
laminar flow regions. Additionally, a well-validated transition model
Figure 4.13. Results from a Lattice Boltzmann solver showing streamwise or empirical approach to determine the boundary-layer transition
velocity component U (m/s) on a horizontal plane at the trailer mid-height location should be used.
versus PIV measurements from a NASA Ames 7-by-10 foot tunnel at Rew of
1.1 million, Heinecke [268].
The influence of CFD on the commercial vehicle industry extends
well beyond external aerodynamics and includes internal flows,
4.6. CFD Method Selections and Use combustion, hydraulics, and thermal management. In each of these
A comparison of the capabilities of five classes of computational areas Reynolds number plays a key role in defining the fluid
methods is shown in Table 4.1. The summary information indicate dynamics and in guiding method application. In 2013 the SAE took a
that the aerodynamics of a complete commercial vehicle cannot be first step in addressing CFD validation and verification with the
represented with either a RANS or URANS method due their publication of recommended practice J2966 “Guidelines for
inability to explicitly model a wide range of boundary layer states, Aerodynamic Assessment of Medium and Heavy Commercial
transition, separation and unsteady flows. Ground Vehicles Using Computational Fluid Dynamics”.

Table 4.1. Features and capabilities of numerical techniques (portions taken Computational simulation methods are based on either the Navier-
from Spalart [239]). Stokes (NS) equations or the lattice Boltzman equation (LBE). Both
methods rely on similar turbulence modeling but have unique
advantages and disadvantages when solving the commercial ground
vehicle problem. Modeling schemes include Direct Numerical
Simulation (DNS) and Reynolds Averaged Navier-Stokes (RANS),
unsteady RANS (URANS), Large-Eddy Simulation (LES), Very-
Large-Eddy Simulation (VLES), and Detached-Eddy Simulation
(DES). However setting Reynolds number in the computational
simulation does not, by itself, result in flow similarity. The latter
requires geometric and kinematic similarity, both of which are
difficult to achieve for a commercial ground vehicle.

When accurate prediction of separated wake flows and the impact on


aerodynamic drag are needed, CFD methods which capture the
turbulent, transient flow structures in separated regions using the
highest possible fidelity should be used. Methods such as VLES or
DES are recommended over RANS so that turbulent flow structures
It is suggested that the use of RANS and URANS should be limited are simulated rather than modeled.
to preliminary design and analysis of vehicle/models components
with attached flow and known boundary layer characteristics. It is suggested that the use of RANS and URANS should be limited
HYBRID and LES methods should be used, with experimentally to preliminary design and analysis of vehicle/models components
determine boundary layer characteristics, for detailed design vehicles/ with attached flow and known boundary layer characteristics.
models, aerodynamic components and evaluation of vehicle HYBRID and LES methods should be used, with experimentally
aerodynamic performance. determined boundary layer characteristics, for detailed design
vehicles/models, aerodynamic components and evaluation of vehicle
aerodynamic performance.
4.7. Summary Findings
The dominant global flow regime for commercial vehicles is defined Computational fluid dynamics is an ever-expanding field of study
as incompressible, unsteady, turbulent, and separated but may have requiring knowledge of fluid dynamics, computational science, and
local regions of compressible, steady, un-separated, and laminar as computer science. Obtaining a relevant solution requires a basic
well as transitional flow. However, commercial vehicles vary greatly understanding of fluid dynamics and computational sciences. To
in size and shape and operate over a broad range of Reynolds number obtain the best available solution requires an ability to make expert
resulting in a diverse range of boundary layer, on-body and off-body judgments in fluid dynamics and computational sciences.4.8.
flow regimes. For example, a vehicle may have areas of laminar and
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 629

4.8. References 252. Spalart P., Jou W., Strelets M. and Allmaras S., “Comments on the
231. Oberkampf, W. and Trucano, T., “Verification and Validation in Feasibility of LES for Wings, and on a Hybrid RANS/LES Approach,”
Computational Fluid Dynamics,” Progress in Aerospace Sciences 38(3): Advances in DNS/LES 1:4-8, 1st AFOSR International Conference on
209-272, SAND2002 - 0529, 2002. DNS and LES, NY, 1997.
232. AIAA-G-077-1998 Guide for the Verification and Validation of 253. Travin A., Shur M., Strelets M. and Spalart P.: “Detached-Eddy
Computational Fluid Dynamics Simulations, ISBN: 978-1-56347-354-8, Simulations Past a Circular Cylinder,” Flow Turbulence and Combustion
1998. 63(1-4):293-313, 1999.
233. ASME V&V 20-2009 Standard for Verification and Validation 254. Bird, R., Stewart, W. and Lightfoot, E., “Transport Phenomena,” New
in Computational Fluid Dynamics and Heat Transfer, ISBN: York, John Wiley & Sons, Inc., New York, 2002.
9780791832097, 2009. 255. von Kármán, T. (1930), “Mechanische Ähnlichkeit und Turbulenz,”
234. Dybenko, J and Savory, E., “An Experimental Investigation of Turbulent Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen,
Boundary Layer Flow Over Surface-Mounted Circular Cavities,” Fachgruppe 1 (Mathematik) 5: 58-76 1930. Also as:“Mechanical
Proceedings of the Institution of Mechanical Engineers Part G, Journal Similitude and Turbulence”, NACA Technical Memorandum 611, 1931.
of Aerospace Engineering 2006(1), CSME Forum, 2006. 256. Uranga, A., “Investigation of Transition to Turbulence at Low Reynolds
235. Gaudet, L. and Winter, K., “Measurements of the Drag of Some Numbers Using Implicit Large Eddy Simulations with a Discontinuous
Characteristic Aircraft Excrescences Immersed in Turbulent Boundary Glerkin Method,” Ph.D. Thesis, MIT, Dept. of Aeronautics and
Layers,” Royal Aircraft Establishment Tech. Memo Aero 1538, 1973. Astronautics, 2011
236. ASME JFE v.130 Celik, B. I., et.al., “Procedure for Estimation and 257. Langtry, R. and Menter, F., “Transition modeling for General CFD
Reporting of Uncertainty Due to Discretization in CFD Applications,” Applications in Aeronautics,” AIAA 2005-522, 43rd AIAA Aerospace
Journal of Fluids Engineering, v. 130, July 2008, doi:10.1115/1.2960953. Sciences Meeting and Exhibit, Reno, NV, January 2005.
237. Pope, S., “Turbulent Flows,” Cambridge University Press, ISBN: 258. Sayadi, T. and Moiun, P., “Predicting Natural Transition Using Large
9780521598866, 2006. Eddy Simulation,” Stanford University, CA, Center for Turbulence
Research, Annual Research Briefs, 2011.
238. Moin, P. and Kim, J., “Tackling Turbulence with Supercomputers,”
Scientific American 276(1):62-68, Jan. 1997. 259. Narasimha, R.: “Modeling the Transitional Boundary Layer,” Institute
for Computer Applications in Science and Engineering, NASA Langley
239. Spalart, P., “Strategies for Turbulence Modeling and Simulations,” Research Center, Hampton, VA, ICASE Report 90-90, also as NASA CR
International Journal of Heat and Fluid Flow 21(3): 252-263, 2000. 187487, 1990.
240. Spalart, P., “Trends in Turbulence Treatments,” AIAA Paper 2000-2306, 260. White, F.M. (1991), “Viscous Fluid Flow,” Second edition, McGraw-
Fluids 2000 Conference and Exhibit, Denver, CO, June 2000. Hill Inc., New York, 1991.
241. DeGraff, D., Webster, R. and Eaton, J., “The Effect of Reynolds Number 261. van Ingen, J., “The eN Method for Transition Prediction, Historical
on Boundary Layer Turbulence,” Experimental Thermal and Fluid Review of Work at TU Delft,” 38th AIAA Fluid Dynamics Conference
Science 18(4):341-346, 1999. and Exhibit, Seattle, WA, AIAA 2008-3830, 2008.
242. Wilcox, D.C., “Turbulence Modelling for CFD,” 2nd edition, DCW 262. Rumsey, C., Slotnick, J., Long, M., Stuever, R. and Wayman, T.,
Industries, inc., ISBN: 0963605151, 1998. “Summary of the First AIAA CFD High-Lift Prediction Workshop,”
243. Menter, F. R., “Zonal Two Equation k-ω Turbulence Models for Journal of Aircraft 48(6):2068-2079, 2011.
Aerodynamic Flows,” AIAA Paper 93-2906, 24th Fluid Dynamics 263. Rumsey, C. and Lee-Rausch, E., “NASA Trapezoidal Wing
Conference, Orlando, FL, July 1993. Computations Including Transition and Advanced Turbulence
244. Nichols, R., “Turbulence Models and Their Application to Complex Modeling,” 30th AIAA Applied Aerodynamics Conference, New Orleans,
Flows,” DoD High Performance Computing Modernization Program LA, AIAA-2012-2843, June 2012. Also in Journal of Aircraft 52(2):496-
(HPCMP) Programming Environment and Training (PET) activities 509, 2014.
through Mississippi State University under terms of Contract No. 264. Eliasson, P., Hanifi, A. and Peng, S., “Influence of Transition on High-
N62306-01-D-7110 Lift Prediction for the NASA Trap Wing Model,” 29th AIAA Applied
245. Bayraktar, I. and Bayraktar, T., “Guidelines for CFD Simulations of Aerodynamics Conference, Honolulu, Hawaii, AIAA 2011-3009, June
Ground Vehicle Aerodynamics,” SAE Technical Paper 2006-01-3544, 2011.
2006, doi:10.4271/2006-01-3544. 265. Shankara, P. and Snyder, D., “Numerical Simulation of High Lift
246. Spalart, R., “Detached-Eddy Simulation,” Annual Review of Fluid Trap Wing Using STAR-CCM+,” 30th AIAA Applied Aerodynamics
Mechanics 41:181-202, doi:10.1146/annurev.fluid.010908.165130, Conference, New Orleans, LA, AIAA-2012-2920, June 2012.
2009. 266. Eichinger, S., “Active Flow Separation Control of Ground
247. Launder B. and Spalding D., “Mathematical Models of Turbulence,” Transportation Vehicle Configurations,” Technical University of Berlin,
Zeitschrift für Angewandte Mathematik und Mechanik (Journal of Faculty V of Mechanical Engineering and Transport Systems, Ph.D.
Applied Mathematics and Mechanics) 53(6): 424ff., 1972. Thesis, Sept 2012.
248. Chen, S. and Doolen, G., “Lattice Boltzmann Method for Fluid Flows,” 267. Horrigan, K., Duncan, B., Keating, A., Gupta, A. et al., “Aerodynamic
Annual Review of Fluid Mechanics 30:329-64, 1998. Simulations of a Generic Tractor-Trailer: Validation and Analysis of
249. Derksen, J., “Lattice-Boltzman Based Large-Eddy Simulations Unsteady Aerodynamics,” SAE Technical Paper 2008-01-2612, 2008,
Applied to Industrial Flow,” Notes in Computer Science 2329:713- doi:10.4271/2008-01-2612.
722, International Conference on Computational Science - ICCS 2002, 268. Heineck, J. T.; Walker, S. M.; and Satran, D., “The Measurement of
Amsterdam, The Netherlands, April 2002. Wake and Gap Flows of the Generic Conventional Truck Model (GCM)
250. Lockard, D., Lou, P. and Singer, B., “Evaluation of the Lattice- Using Three-Component PIV,” The Aerodynamics of Heavy Vehicles:
Boltzmann Equation Solver PowerFLOW for Aerodynamic Trucks, Buses and Trains, 19:173-184, 2004.
Applications,” NASA CR 2000-210550, 2000
251. Alfonsi, G., “Reynolds-Averaged Navier-Stokes Equations for
Turbulence Modeling,” Applied Mechanics Reviews, 62(4):040802,
doi:10.1115/1.3124648, 2009.
630 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

CHAPTER 5. EXPERIMENTAL TESTING test articles. When comparing results obtained from different
methods, different test articles and/or different facilities the precision
Experimental test methods employed to evaluate the aerodynamics of
of the comparison requires an analysis of the individual uncertainty
commercial vehicles are; wind tunnel test, coastdown, constant speed,
values from each test, as discussed in references 280, 281, 282, 283.
and fuel consumption testing. Wind tunnel testing is the only
experimental method that obtains aerodynamic force and moment data
All experimental test methods include assumptions and constraints
[3, 14, 62, 95]. The second data type is road-load values that are
that impact the repeatability of the data. These assumptions can be
generated from both coastdown [19, 131, 207, 269, 270] and constant
different for each test methodology. SAE J1263 [269] coastdown
speed test [271, 272]. Note that a constant speed test has not been
testing allows a variation in atmospheric effects during the test process
developed at this time and work is in progress in North America,
and assumes the aerodynamics of the vehicle is independent of vehicle
Europe and Japan. The third and final method is fuel economy testing
speed. In comparison SAE J2263 [270] coastdown procedure assumes
that generates fuel use data through the operational testing of the
the use of a vehicle-mounted anemometer has minimal effect on the
vehicle [94, 134, 135]. Associated with each test methods is a range of
aerodynamics of the vehicle; see Figure 5.1 [284]. Variations in
assumptions and sources of errors that impact the data quality.
atmospheric conditions, vehicle speed and interfering flow fields are
Recognizing the limitations of each test method and facility used for
variables with known Reynolds number sensitivity.
the test is critical to establishing a credible data set. To guide
commercial vehicle testing the SAE has established recommended
practice J1252 [64] for wind tunnel testing and for fuel consumption
testing the two primary methods are J1321 [134] and J1526 [135]. At
this time the SAE is actively developing coastdown and constant speed
test methods for commercial vehicles with an expected publication date
in 2016. To address immediate needs the commercial vehicle industry
has adopted various coastdown procedures [19, 273] of which several
are based on automobile coastdown methods SAE J1263 [270] and
J2263 [271]. Representative experimental test data and information can
be found in references, 13, 31, 41, 64, 66, 117, 136, 144, 145, 147, 149,
150, 156, 164, and 272, 273, 274, 275, 276, 277, 278, 279.

In an effort to maximize the benefit from each test there is a continual


effort to expand the use of the results obtained from each method. For
example, aerodynamic drag values are routinely estimated from road
load data and fuel economy estimates have been derived from wind
a. Photograph of vehicle used in coastdown testing
tunnel aerodynamic data. However, regardless of the test method,
Reynolds number effects will influence the measured values and thus
will affect any result derived from the measured value. Wind tunnel and
coastdown tests are particularly susceptible while constant speed and
fuel economy testing sensitivities are typically limited to atmospheric
and interference phenomena such as cross winds and turbulence.

5.1. Experimental Accuracy, Precision and Uncertainty


The accuracy (bias) of any commercial vehicle experimental test
cannot be determined due to the lack of an established reference value.
b. Time average computational results depicting anemometer flow
A reference value is typically established by repeat measurements of interference.
a standard item that is traceable by a recognized organization such as
National Institute of Standards and Technology (NIST) or Figure 5.1. Representative boom anemometer installation and flow
International Organization for Standardization (ISO). In contrast, interference [284].
precision (repeatability) of a test result can be determined and
Wind tunnel test assumptions are numerous and relate to test flow
expressed as an uncertainty range that characterizes the possible
characteristics, model support method, floor/road simulations, yaw
values of the measured result within a specific confidence interval.
simulations, unsteady forces and various facility interference effects.
Precision is the ability to get similar results through repeated testing
Assumption related to test article geometry is a constant area of
of the same test article in the same facility using the same method.
concern. Tests of “as-built” vs. “as designed” highlight tolerance
Understanding the precision of a test is required to assess data quality
differences from manufacturing and operation. Scale modeling
and to determine test relevance. To address this critical need the SAE
presents additional opportunities for deviation from the “as-built”
has incorporated uncertainty analysis into the J1252 [64], J1321 [134]
full-scale vehicle. Failing to satisfy geometric similarity criteria will
and J1526 [135] test methods. Note, repeatability is not associated
violate flow similarity criteria and thereby Reynolds number
with the comparison of results from different; facilities, methods or
matching will not be satisfied.
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 631

All of these assumptions affect the freestream flow and/or the data values represent different test conditions it is known that
on-body flow characteristics, which are governed by the Reynolds aerodynamic drag varies directly with increasing yaw angle but may
number. Comparing data obtained at different freestream and/or local vary directly or inversely with increasing Reynolds number. Neither
Reynolds number values and boundary layer conditions contribute to of these effects is accounted for in the calculation of the WAD value.
variability in CD prediction and would not be expected to generate
similar results, see Chapter 2. Obtaining similar results from different The data of figure 5.2 highlight the importance of performing
methods or different test articles with different test conditions is aerodynamic testing at Reynolds number values in which the
coincidental and not validation. boundary layer characteristics and aerodynamic forces are similar to
those of the full-scale vehicle at operational conditions. This requires
An example of possible variability between methods is seen in data the measurement of the boundary layer characteristics of the
taken from the Environmental Protection Agency (EPA) Greenhouse full-scale vehicle at operational conditions. If tests do not match the
Gas (GHG) Regulatory Impact Analysis report issued in 2011 [284], boundary layer characteristics of the full-scale then testing should
see figure 5.2. The chart shows results for sub-scale and full-scale utilize methods to modify the boundary layer characteristics to match
wind tunnel tests and coastdown test. Each test was executed with full-scale conditions.
different test articles that were fabricated based on identical geometry
definitions. The coast down data was obtained over a width-based
Reynolds number range of 5.5×106 to 1.0×106 and the wind tunnel
5.2. Wind Tunnel Testing
test data was obtained at width-based Reynolds number of The importance of Reynolds number similarity to wind tunnel testing
approximately 4.5×106 and 1.0×106, for the full-scale and sub-scale cannot be understated and is based on the fundamental requirement
date respectively. The subject EPA report does not document the that flow similarity must be satisfied for a wind tunnel test to produce
boundary layer state of the operational vehicle or indicate that the results that represent the operational commercial vehicle. The
boundary layer for either the coast down or the wind tunnel tests Reynolds number requirement is clearly defined in SAE standard
replicated the characteristics of the boundary layer state of the wind tunnel test documents J2084 [285] and J1252 [64].
operational vehicle. The data is presented without an uncertainty
assessment of the individual test values and without an analysis of the From J2084 [285]:
combined uncertainty of the difference between test values. The lack
of an uncertainty analysis limits the validity of a comparison between “A general aerodynamic requirement to ensure reliable wind-tunnel
the individual test results. measurements is that correct Reynolds numbers have to be simulated”.

Assuming that the combined uncertainty of the differences between SAE recommended practice J1252 [64] provides guidelines for wind
the test results is much less than the differences between the tunnel testing of commercial vehicles. Recognizing both the
measured values then we can make a few general observations. The importance of Reynolds number to flow similarity and the limited
data of figure 5.2 suggest that the difference in drag may be attributed number of wind tunnels capable of testing full size commercial
to variations in; modeling of geometric details, boundary layer state vehicles or testing at full scale Reynolds numbers the SAE J1252
and characteristics, and Reynolds number effects. recommended practice provides additional information for testing at
sub-scale Reynolds number, as noted in the following two statements.

From J1252 [64]:

“Tests should be performed to determine the effect of vehicle


Reynolds number (Rew) on the measured forces and moments over the
range of yaw angles dictated by the test requirements. When the
Reynolds number effects have been determined, the remainder of the
test program should be performed at a Reynolds number above which
the force and moment coefficients are essentially invariant for the
yaw angles investigated.”

“In the case where the overall Reynolds number of the model may
Figure 5.2. Comparison of data for coastdown, sub-scale wind tunnel and
not match the on-road vehicle, boundary layer characteristics
full-scale wind tunnel tests. This figure is adapted from the data presented in
measured on the model should match the boundary layer
reference 284.
characteristics at the same locations on the full scale vehicle. The
Another limitation of the comparison shown in figure 5.2 is the use of criteria may be satisfied with or without boundary layer roughness
the wind-averaged drag (WAD) value for comparing wind tunnel test elements on the model. As a minimum the turbulent boundary layer
data to coastdown test data. A discussion and explanation of the WAD profiles of the tested model should match the road vehicle. This is
value can be found in SAE J1252 [64]. Beyond the fact that the two especially important at free edges.”
632 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

5.2.1. Reynolds Number Similarity The need for similarity is related to use of the test data. If the intent is
Simply stated, Reynolds number similarity is achieved when the to perform exploratory testing of generic aerodynamic concepts
geometry, test environment and all Reynolds number dependent forces, Reynolds number similarity may not be a requirement. However if
moments, pressures, boundary layer characteristics as well as on and the generic concept data is to be related/compared to any other data
off-body flow characteristics is equivalent for each test result being set or to an operational vehicle then similarity criteria must be
compared. The most direct and most difficult approach is to exactly satisfied between data sets. Additionally a comparison may not
replicate the geometry, Reynolds number and test environment. achieve similarity for the complete vehicle but similarity may be
satisfied for a component or limited region of the vehicle. An
Determining the Reynolds number value to achieve similarity for a example would be an investigation of a base treatment such as a
wind tunnel test can be difficult. The SAE J1252 [64] guideline, is boattail. If the Reynolds number dependent forces, moments, local
based on historical data for heavy combination vehicles, requires pressures, boundary layer characteristics as well as on and off-body
Reynolds number sweeps at constant yaw angles between 0 degree flow characteristics upstream of the boattail are equivalent then the
and the maximum yaw angle to establish the Reynolds number above Reynolds number dependent forces, moments, local pressures,
which the force and moment coefficients are “essentially constant” boundary layer characteristics as well as on and off-body flow
(i.e. transcritical state), see representative curves in figures 3.1, 3.13, characteristics of the boattail would be equivalent.
3.14, 3.15, 3.16, 3.17. The minimum test Reynolds number is defined
as the largest Reynolds number value obtained from the fixed The path to Reynolds number similarity is complex and dependent on
yaw-angle Reynolds number sweeps. For example the data of figure many factors. The following guidelines may be used to determine the
3.14 show that the Reynolds number sweep at 10 degrees yaw angle wind tunnel test Reynolds number.
results in a minimum test Reynolds number of 5.0 million. This
Reynolds number value exceeds those obtained for the 0 degree and 5 For wind tunnel comparison to operational vehicles:
degree yaw angle Reynolds number sweeps as a result the minimum
test Reynolds number to satisfy both J1252 and J2084 is 5.0 million. • If a vehicle operates in the subcritical or transitional range the
Figures 3.15, 3.16, 3.17 show data for a yaw angle of 0 degree that wind tunnel test should be performed at the vehicle's operational
indicates a minimum test Reynolds number between 3.0 and 4.0 Reynolds number.
million. This 0 degree yaw angle based minimum Reynolds number • If a vehicle operates in the transcritical range the J1252 criteria
range is consistent with the data of Figure 3.14. Note: the data should be modified in which the “essentially constant” criteria
presented in figures 3.13, 3.14, 3.15, 3.16, 3.17 is for heavy are defined as the point that the slope (Str) of the CD with
combination commercial vehicles. Reynolds number becomes and remains less than 0.005, see Eq.
5.1. The minimum test Reynolds number shall be ten percent
A review of the CD versus Reynolds number data of Chapter 3 show greater than the minimum transcritical Reynolds number value
that the variation in the subcritical, transitional and transcritical that satisfies the slope criteria.
ranges are vehicle dependent and vary significantly between vehicles
within a class and between classes of commercial vehicles. In
general, the subcritical and transcritical ranges vary in a more linear (5.1)
fashion than the transitional range. The subcritical range may be a
constant value or could display small to moderate increases or For wind tunnel comparison to operational vehicle component data:
decreases with increasing Reynolds number. The transitional curve
segment may show a steep or gradual decrease or increase in drag or • If a vehicle operates in the subcritical or transitional range the
other nonlinear trends between the subcritical and transcritical range. wind tunnel test shall be performed at the vehicle's operational
In all cases the transcritical range is defined as the condition in which Reynolds number.
the aerodynamics of the vehicle become invariant with all further • If a vehicle operates in the transcritical Reynolds number range
increases in Reynolds number. the modified J1252 criteria is preferred. An alternate local
criteria may be used requiring Reynolds number dependent
These observations suggest that the use of the transcritical Reynolds forces, moments, local pressures, boundary layer characteristics
number value proposed in J1252 [64] may not be appropriate for all as well as on and off-body flow characteristics upstream of the
current and future vehicle types and their operational speed. For device or localized area of each data set to be equivalent.
example a single unit delivery vehicle with an average speed of 35 mph
has an operational Reynolds number of 2.7 million. This value For comparison between wind tunnel test:
corresponds to the transitional range for this class of vehicle. The
characteristic large variation in aerodynamics with small changes in • If either test is performed at subcritical or transitional range both
Reynolds present in the transitional range dictates that Reynolds number wind tunnel test shall be performed on the same geometry, at the
similarity can only be achieved with a Reynolds number of 2.7 million. same Reynolds number and under equivalent test conditions.
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 633

• If the tests are performed in the transcritical range the modified


J1252 criteria shall be used to evaluate both test data sets to
determine the minimum allowable test Reynolds number for
each test. Both tests shall be performed at the same Reynolds
number value and under equivalent test conditions.

5.2.2. Wind Tunnel Facilities


Commercial vehicle wind tunnel test objectives typically support new
vehicle development or improvements to existing vehicles. The latter
of these two test types may take many forms in which the primary
test article may vary from a generic shaped body to an existing
vehicle design. Figure 5.3 depicts a range commercial vehicles test
c. 10 percent scale generic vehicle in the Georgia Tech Research Institute
articles and wind tunnel facilities. Information on wind tunnels that
(GTRI) low speed wind tunnel with closed test section [288]
may be used to support commercial vehicle aerodynamic studies can
be found in the Appendix.

The wind tunnel facilities shown in figure 5.3a, 5.3b and 5.3c have
closed test sections while the wind tunnel shown in figure 5d has an
open test section as graphically depicted in figure 5.4. The type of test
section impacts model size and range of yaw angles that will affect
Reynolds number similarity. As shown in figure 5.5 an open test
section can accept a model that is sixty percent larger than a closed
test section. As shown in figure 5.5 model size is limited by the
blockage constraint applied at maximum yaw angle.

d. 25 percent scale generic vehicle in NASA Full Scale wind tunnel with open
test section.

Figure 5.3. Photographs of operational and generic vehicles in various wind


tunnel facilities.

a. Operational tractor in the Freightliner wind tunnel [286]

Figure 5.4. Graphic depicting open and closed test sections.


b. Operational vehicle in the National Research Council (NRC) of Canada 9
meter wind tunnel with closed test section [287]
634 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

vehicle’s leading edge to no more than 10 percent of the vehicles


minimum ground clearance. SAE J1252 ground simulation guideline
is to use a static ground plane with a boundary layer removal system.

Figure 5.5. Graphic depicting maximum blockage for open and closed test
sections.

Additionally the test objective may be to obtain force and moment


data, on and off-body flow information, acoustic signature, engine and
Figure 5.6. Sketch of various wind-tunnel floor modeling methods.
vehicle cooling performance [289] or climatic effects [290]. Details
related to the various test objectives may be found in J1252 [63]. In all From J1252: “The large yaw sweeps, non-constant ground clearance
cases the test article may be sub or full-scale and the data obtained and Reynolds number dependence of a commercial road vehicle
will vary with facility selection. Data variability is related to makes this recommendation conservative.”
management of the floor and model boundary layers, turbulence levels
in the free stream, model support method/corrections, blockage, and An alternative is the use of a raised ground plane that is offset from
boundary interference and corrections. Minimizing data variability the tunnel floor. Both of these methods are shown in the middle
between facilities and data uncertainty within a test requires that each column of figure 5.6 and represent the methods used in the wind
test satisfy the flow similarity and thus Reynolds number similarity tunnels shown in figures 5.3b, c, d.
requirements discussed in section 5.2.1 and Chapter 2.
It has been discussed in the literature that a rolling floor/road concept
5.2.3. Floor/Road Simulation with rotating wheels would provide a more accurate simulation for
A vehicle in operation is moving through air and a boundary layer commercial vehicle testing, however published results have not
develops on all wetted surface including the vehicle lower surface. In demonstrated this claim. The automobile and racing industries prefer
contrast, for wind tunnel tests the air is flowing through the test this method due to the vehicle's low ground clearances, low sensitivity
section and a boundary layer will develop on the floor of the test to yaw, and increased importance of lift. In comparison commercial
section as well all other surfaces located in and comprising the test vehicles have large ground clearances, a high sensitivity to yaw (see
section, including the floor under the vehicle and on the vehicle lower sub section 5.2.5) and lift is of minimal interest. Of the three factors
surface. The fundamental objective of wind tunnel floor simulation yaw is the most significant when discussing the use of a rolling floor
for commercial vehicle testing is to model the boundary layer for commercial vehicle tests. Implementation of the rolling floor
characteristics present on the road under and in proximity to an method varies greatly in the number of rolling floor segments [291,
operating vehicle [14, 15, 40]. A series of sketches representing 292, 293, 294] and in the relationship between rolling floor and model
various techniques are presented in figure 5.6. The simulation with increasing yaw angle. Yaw angle is implemented in two ways, the
strategy is to minimize the momentum deficit in the floor boundary floor fixed at 0 degree yaw angle or the floor and the test article yaw
layer. As shown in the figure road simulation can be accomplished together [294]. Both yaw methods fail to model the boundary layer
passively by raising the test article out of the floor boundary layer as characteristics present on the road under and in proximity to the
depicted in the standoff and mirroring concepts in the left column of operating vehicle and create a significant discontinuity at the front and
figure 5.6 or by removing or adding mass or energy to the floor windward edges of the yawed moving floor or the test article.
boundary layer, see middle and right hand columns of figure 5.6.
Current practice within the industry is focused on one or a 5.2.4. Model Support
combination of active methods. A model support system performs two functions: 1) positions the
model in the test section and 2) either provides the mechanical
A detailed discussion of these ground simulation options for connection between an internal balance and the facility structure or
minimizing the interference between test article Reynolds number between the model and an external balance [14, 40, 64]. The method
effects and ground simulation Reynolds number effects can be found of model support is a function of; model design, data requirements,
in references 14, 40, 64 and 285. SAE recommended practice J1252 aerodynamic loads, test section type, ground simulation technique
[64] guidance is to limit the height of the ground boundary layer and force balance location. Common to all installations is the desire
displacement thickness relative to the ground clearance at the to eliminate the air gap between the tires and the ground plane (floor)
[1, 14, 40, 64]. In performing these tasks the model support should
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 635

not impact the aerodynamic test results. For an external balance there Suggested model support design criteria to minimize aerodynamic
are three areas of concern; 1) loads on the model support, 2) the interference are:
model support altering the flow on the model and 3) model altering
the flow on the model support. For an internal balance there are two 1. Surface area exposed to high-velocity/free-stream flow < 10
areas of concern; 1) the model support altering the flow on the model percent of model cross section
and 2) the flow penetrating the metric seal between the model support 2. Axial projected area < 1 percent of model cross section
and the model. All five areas of concern will vary with Reynolds 3. Reynolds number sensitivity = constant
number. To minimize these effects various model support systems are
4. Yaw sensitivity = constant
used and several examples are depicted in figure 5.7. Sub-scale
vehicle testing may use various methods due to the low weight of the
SAE J1252 further recommends a test program to determine the
model however tests of full size vehicles will typically use the wheel
aerodynamic interference corrections and Reynolds number
pad support system shown in the upper right of figure 5.7. The wheel
sensitivities. This step is critical if the geometric variation is either in
pad system is used with an external balance whereas the other three
close proximity to or down stream of the model support structure.
systems depicted in figure 5.7 may use either an external or internal
balance system.
5.2.5. Yaw Sweep
Atmospheric effects in concert with traffic and both natural and
manmade road-side structures result in an on-road operating
environment of significant cross winds (yaw) and varying levels of
turbulence [192]. The sensitivity of commercial vehicles to yaw is
greatly pronounced compared to that for automobiles. This sensitivity
is clearly evident in a comparison of drag with yaw angle for various
vehicle classes shown in figure 5.8 [64]. The chart shows that drag
for commercial vehicles increases with yaw at a rate six to ten times
greater than seen for an passenger car. These data clearly highlight
the importance of yaw data in wind tunnel testing.

Yaw testing in tunnels is constrained by the facility, test support


hardware and test vehicle dimensions. Operational yaw angle results
Figure 5.7. Sketch of various model support methods. from the vector summation of the vehicle forward speed and the wind
speed and angle relative to the vehicle centerline.
A concern present with all support systems is the interference with
the model and impact on the resultant forces and moments. This
interference will, at a minimum alter the flow on the model in close
proximity to the intersection of the support system with the model as
well as all loads downstream of the intersection point [295, 296, 297,
298]. Of particular concern is modification to the either the floor or
model boundary layer resulting from support induced pressure
disturbances and flow separation.

Aerodynamic interference assessment studies have shown that the


effects may be significant and should be minimized and/or corrected
[295, 296, 297, 298]. However, the complexity of the aerodynamic
interference effects and their variability with changes in vehicle
geometry, Reynolds number, and yaw conditions makes it highly
unlikely that a simple correction can be derived to account for all test
conditions and test models. The problem is further compounded if the
desired measurement resolution is extremely small or the geometric
variation is in close proximity to the model support structure.
Figure 5.8. Variation in drag for various vehicle classes [64].
As outlined in SAE 1252 are guidelines for minimizing the
aerodynamic interference of the support system. First, minimize the The resultant of these two vectors is the yaw angle and an effective air
surface area of the support system exposed to the high velocity speed of the vehicle that will differ from the vehicle speed relative to
freestream. Second, minimize the axial projected area of the support the road surface. The effective air speed may be less or greater than the
system. The third and fourth criteria are to eliminate the Reynolds vehicle road speed. In the wind tunnel the yaw angle is created by a
number and yaw aerodynamic sensitivity of the support system. rotation of the model in a constant velocity air stream. The result is a
yaw angle and a representative vehicle speed (ie velocity vector aligned
636 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

with the model centerline) that is lower than the tunnel flow velocity. SAE J1252 guidelines related to blockage are summarized below.
Further, the simplified wind tunnel representation of cross wind is
acceptable if the vehicle's operational Reynolds number is in the • Area blockage should be calculated based on the projected
transcritical range and the test is performed at Reynolds number values frontal area of the vehicle at the maximum tested yaw angle. In
that satisfy Reynolds number similarity requirements of J1252. For addition, vehicle length and proximity to the beginning and end
example an analysis of the Storms [55] data of figure 3.13 and 3.14 of the test section are important considerations.
identifies that the minimum test Reynolds number is 5.0 million. A test • For closed wall test section the test blockage ratio be limited
at 1.0 million Reynolds number would introduce an error in drag of 7.0 to 5 percent at the maximum yaw angle used, and that the
percent, 5.0 percent and 2.5 percent at yaw angles of 10 degree, 5 magnitude of the corrections be limited to a 10 percent change
degree and 0 degree, respectively. Additional Reynolds number based from the overall measured drag coefficient.
errors may be introduced in the wind tunnel if the change in Reynolds • For open jet test section the test blockage ratio be limited
number due to head wind or tail wind is not modeled. to 10 percent at the maximum yaw angle used and that
corrections be limited to a 10 percent change in the overall
5.2.6. Additional Factors measured drag coefficient.
From reference 156; “In general, the aim of wind-tunnel tests is to
make measurements of aerodynamic quantities under strictly Another area of concern, that is not typically studied and corrected, is
controlled and defined conditions in such a way that, despite the the on-model boundary layer characteristics that will impact total
presence of the tunnel walls, the data can be applied to unconstrained forces and moments as well all time dependent flow features. The use
flow. The existence of a free-air flow which is equivalent to that in the of boundary layer assessment methods such as a boundary layer
tunnel is the fundamental assumption underlying the entire survey rake [195, 306, 307] should be used to determine boundary
framework of the theory and practice of wind tunnel wall constraint.” layer state, thickness and transition locations, see figure 5.9. These
measurements can be used with various flow visualization methods
This statement can be expanded beyond wall corrections [279, 299] [14, 88, 149, 277, 309], such as fluorescent oil [308], and forced
and include; boundary layer transition techniques as shown in figure 5.10 [195] to
satisfy Reynolds simulation criteria as outlined in J1252 and
1. model support [294, 295, 296, 297, 298], discussed previously in section 5.2.

2. floor modeling [ 291, 292, 293, 294],


3. freestream turbulence [96, 152, 167, 300],
4. yaw modeling [55, 64],
5. model size [226, 279, 301],
6. boundary layer state [44, 147, 148, 195, 302] and
7. test section type [24, 230].

To predict operational vehicle performance from wind tunnel test


results require that all corrections are applied and all Reynolds
number sensitivities are identified and modeled. In addition, the test
article must be geometrically and kinematically scaled to a very high
degree of detail. Wind tunnel tests typically do not satisfy geometric,
kinematic or Reynolds number similarity for;

1. flow stream related to the underhood region including thermal


characteristics [46, 303, 304, 305],
2. wheels/tire flow and mechanical dynamics [38, 215, 224, 306],
and
3. on-model local and global boundary layer characteristics [117,
194, 241].

These items as well as others require specialized treatment that may a. Conventional boundary layer rake [308].
impact the previous listed corrections to the wind tunnel data set.
Fig 5.9. Photographs of boundary layer rake.

An area of concern is that traditional correction for blockage (i.e.


model size) results in an adjustment to the dynamic pressure [64,
226]. This correction may impact the model flow characteristics that
in turn would impact the Reynolds number sensitivities of the model.
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 637

5.3. Road Load Testing


Road load data is obtained from either coastdown methods [19, 131]
or constant speed testing [273, 274]. While coastdown testing has been
considered a testing option for more than half a century [18, 309, 310,
311, 312, 313, 314, 315, 316, 317, 318, 319, 320] the latter method is
relatively new and as such there is minimal published information
available. Coastdown testing obtains road load data by measuring the
rate of deceleration of the vehicle as it decelerates from a high speed
to a low speed. The constant speed test obtains road load data by
measuring the torque at the wheels as the vehicle is operated at a
different constant speed conditions. Both methods assign estimated
values to mechanical contributors, assume a variability, or
invariability, for other values or attempt to make uncontrolled factors
irrelevant by conducting tests in controlled conditions.

The majority of the coastdown methods in use for commercial vehicles


are based on either the SAE J1263 [269] or SAE J2263 [270]
recommended practice, such as the EPA GHG regulation CFR
§1066.310 Coastdown Procedures for Heavy-Duty Vehicles. Both the
J1263 and J2263 were developed for automobiles and lack constraints
necessary to obtain results for bluff body commercial vehicles. All of
the SAE based methods allow coastdown speeds below 15 mph. An
alternative approach is the coastdown method employed by NASA in
b. Boundary layer rake installed on scale model. the 1970s [320] to investigate commercial vehicle drag reduction
technology. The NASA methods specified testing from 65 to 35 mph
Fig 5.9 (cont). Photographs of boundary layer rake. with ambient winds not to exceed 5 mph. NASA's increase in the low
end of the allowed coastdown speed to 35mph was partly attributed to
a concern for low Reynolds number effects contaminating the data. The
constant speed test obtains data at low speed but this data is only used
to evaluate rolling resistance and is not used to derive the aerodynamic
load and therefore low Reynolds number effects are not a concern.

Both Buckley [319] and Walter [321] acknowledge the low


Reynolds number test condition. While Buckley in 1995 evaluated
available wind tunnel data at the time and concluded that it was
acceptable to assume drag coefficient is independent of speed,
Walter in 2001 seemed less certain and pointed out the fixed data
collection time interval weighted test results to speeds 30mph and
below. He notes that raising the low-end speed to 40mph had a
a. Without forced transition negligible impact on the test result.

Equally important is the ambient wind allowance for each method. It


is well understood that a vehicle's drag coefficient is yaw dependent,
so to perform coastdown testing in the presence of wind guarantees a
continuously variable CD throughout the vehicle's deceleration. To
assess Reynolds number effects in coastdown testing a comparison of
the coastdown effective speed and yaw angle due to wind allowance
on road load and aerodynamic drag for the J1263, J2263, EPA GHG,
and NASA test methods is presented in tables 5.1, 5.2, 5.3, 5.4. These
data are referenced to a commercial vehicle operating under the
following conditions;

1. Elevation : 0 ft (m)
b. With forced transition. 2. Temperature : 58.3 °F (14.6 °C)
3. Average US wind speed : 7mph (11.3 kph) [64]
Fig 5.10. Fluorescent oil flow images on generic bluff body model [195].
638 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

4. Vehicle speed : 60 mph (96.6 kmh) criteria varies directly with yaw angle. The allowance of yaw angles
5. Reference length : 8.5 ft. (2.6 m) that exceed the reference conditions of a typical vehicle would
6. Reynolds number: 4.9 million introduce significant variability to the data set that would not reflect
the aerodynamics of the operational vehicle.
Table 5.1 shows the range of vehicle Reynolds numbers for the range
of coastdown speeds, adjusted by the allowable ambient wind The following text is brought forward from Chapter 2. At width-based
conditions aligned parallel to the vehicle centerline. Table 5.2 shows Reynolds number for a commercial vehicle at highway speeds the
the approximate range of maximum yaw angles that may be boundary layer on the forward portion of the vehicle or vehicle
experienced by the vehicle during a test. The yaw angles listed are component, such as the tractor of a tractor-trailer, will have small but
based on the ambient wind aligned perpendicular to the vehicle measurable percentages of both laminar and turbulent boundary layers,
centerline. Table 5.3 shows the extent of laminar boundary layer however, the boundary layer on the trailer of a combination vehicle will
based on the Reynolds number range shown in Table 5.1. Note the be predominately turbulent. The extent of laminar flow and the thickness
laminar extent is based on a transition Reynolds number of 1.0 of the boundary layer will increase with decreasing Reynolds number.
million. Table 5.4 shows estimated change in vehicle aerodynamics For example, a flat plate analysis indicates that at 1.0 million,
due to allowed variation in coastdown speed and wind conditions. representing a full scale vehicle operating at low speed or testing a
The supporting reference for each value is noted in the table. sub-scale model, laminar boundary layer flows may be present over a
significant portion of the front of the vehicle and the development of a
Comparing the values listed in tables 5.1, 5.2, 5.3 to those for the fully turbulent boundary layer will occur significantly downstream
reference vehicle under operational conditions show the NASA test compared to the full-scale vehicle operating at highway speeds. The
method is most consistent with the Reynolds number range, yaw result is a change in separation characteristics and vehicle drag.
angles and extent of laminar boundary layer flow. J1263, J263 and the
EPA GHG methods allow testing that falls well outside the The values shown in table 5.3 clearly indicate that data obtained at
operational characteristics of an on-highway commercial vehicle. vehicle speeds below 35 mph will have significant laminar flow that
would impact the vehicle aerodynamics and more importantly would
The Reynolds number range for each test method shown in Table alter the aerodynamic characteristics of various vehicle components
5.1 extends below the 3.0 million minimum Reynolds number that have a streamwise length less than the width of the vehicle. As
value required for flow similarity identified by Wood [2], Storms mentioned for tables 5.1 and 5.2 only the NASA method minimizes
[55], Leuschen [65] and McArthur [66], the NASA test criteria this unintended effect.
show a 2.5 million lower bound that would result in a minimal
impact to data quality. Insight into the potential impact on test quality/relevance resulting
from each method's test constraints is presented in Table 5.4. The
In contrast to the Reynolds number values of table 5.1, the yaw angle table presents three figures of merit;
values shown in table 5.2 are greater than the reference operational
conditions, with the exception of the NASA method. The allowable 1. Percent change in CD due to the range in Reynolds numbers
test conditions for SAE J1263 and SAE J2263 allow for yaw angles allowed noted in Table 5.1,
that exceed the reference values over the full coastdown range. The 2. Percent change in CD for the range of maximum yaw angles
EPA method satisfies the operational conditions at the high-speed end noted in Table 5.2 over the range of Reynolds numbers noted in
(see Lower Yaw Angle value) of the coast down test but exceeds the Table 5.1, and
reference value for all speeds below 45 mph. As noted in references 3. Potential for boundary layer separation due to the increase in
2, 55 and 64, Reynolds number effects and satisfying the similarity laminar flow noted in Table 5.3.

Table 5.1. Range of Reynolds numbers based on vehicle speed and maximum allowed wind velocity applied as tail or head wind.
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 639

Table 5.2. Range of maximum yaw angles based on vehicle speed and maximum allowed wind velocity applied perpendicular to vehicle axis.

Table 5.3. Range of the extent of laminar boundary layer based on Reynolds number range shown in Table 5.1.

Table 5.4. Estimated change in vehicle drag due to allowed variation in test speed and wind conditions.

An underlying concern is, if the intent of the test is to characterize the vehicles operating in subcritical and transitional Reynolds number
operational performance of the vehicle being tested then it would be state would be highly time-dependent and sensitive to small changes
necessary to know the Reynolds number state for the subject vehicle in speed, wind and temperature. These sensitivities may be present
and to tailor the coastdown and constant speed test conditions to fall for the complete vehicle or the vehicle component under
within the criteria for the specific Reynolds number state. This investigation. Another area of concern is that both SAE J1321 [134]
requires that the Reynolds number state of the test vehicle at and J1526 [135] methods allow for an average wind speed of 12 mph,
operational conditions of interest be determined prior to testing, see this far exceeds the 7mph average wind speed in the continental US
Chapter 3. The following guidance is offered based on the vehicle [64]. This wind allowance would introduce Reynolds number yaw
operational conditions of interest; 1) for subcritical Reynolds number sensitivities, due to changes in flow separation that are not typically
range operations the test shall not be constrained by Reynolds found on an operational vehicle. This Reynolds number yaw
number 2) for transitional Reynolds number range operations a test sensitivity would have a greater impact for SAE J1526 tests, because
shall not be performed, and 3) for transcritical Reynolds number the vehicles are not identical.
range operations the test minimum width-based Reynolds number
shall be greater than 3 million. Both SAE J1321 and SAE J1526 provide a relative difference between
two vehicles operated at the same time over the same test route or
track. The difference between the methods is that J1321 is used to
5.4. Fuel Economy Testing evaluate components whereas J1526 is used to evaluate complete
Unlike wind tunnel testing and road load testing the Reynolds vehicles. To minimize Reynolds number effects, J1321 satisfies the
number sensitivity related to fuel economy or fuel use testing [134, geometric similarity criteria by requiring the two test vehicles to be
135], is less significant if the following criteria are satisfied; 1) test identical with the exception of the component being tested. To
conditions match the operational conditions of interest and 2) the minimize Reynolds number effects, J1526 testing should set the test
vehicles at operational Reynolds number state is known prior to the conditions to ensure that both vehicles are operating in the same
test. This second criteria is critical when interpreting test results that Reynolds number state. As vehicles become more aerodynamic the
may exhibit large run-to-run and segment-to-segment variability importance of both criteria will increase for both SAE test methods.
resulting from an inability to maintain Reynolds number similarity
during the test. The on- and off-body flow and aerodynamics for
640 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

Both fuel use methods require circular or oval shaped test tracks that Consideration of crosswind conditions further accentuates the
results in large portions of the test being conducted with the vehicle Reynolds number sensitivity [2]. Testing in zero ambient wind
skewed to the direction of travel, this effect is magnified for a conditions is very rare, and even low speed ambient winds
combination vehicle, see figure 5.11. An area of concern raised by significantly impact drag.
Mihelic in 2013 [228] is the use of oval or circular test tracks can
introduce Reynolds number effects not found in normal operation Mihelic [228] quantified the misalignment impact with computational
resulting from the effective change (ie misalignment) in vehicle simulations of a representative combination vehicle operating at zero
geometry that is not corrected for in either fuel use test method. Any yaw angle on a curved track, see figure 5.13. The simulation results
vehicle operating on a curved section of road will present a different of figure 5.13 show the outward facing trailer side edge, see right side
geometry to the air than that on a straight road section, see figure image, is exposed to the flow leaving the trailing edge of the outward
5.12. Mihelic states that the component misalignment accentuates the facing tractor aft side fairing. The flow is seen to expand around the
Reynolds number sensitivity of the vehicle due to; the increase in trailer side edge generating an aerodynamic thrust force. The inward
effective vehicle cross section area, varying effective yaw angle along facing side of the vehicle shows the trailer side edge is shielded by
the vehicle length, and the change in interference flows between the tractor and does not exhibit the expected flow expansion around
components. The result is a failure to attain geometric similarity the trailer side edge. The result of the simulation was an 18 percent
between the test condition and the operational condition being increase in vehicle drag compared to a vehicle without misalignment.
simulated by the test.

Figure 5.13. Computational simulation of surface pressures due to


misalignment at yaw = 0 degree. [228]

The asymmetric flow would produce Reynolds number dependent flow


separation, shear flows and vortex structures that would alter the flow
conditions and aerodynamic loading downstream of the asymmetry.
These flow features would impact the performance of all trailer
Figure 5.11. Photograph showing the tractor/trailer misalignment in a track mounted aerodynamic devices as well as that of the tractor cab fairings.
curve. [228]
The magnitude of these effects and the influence of atmospheric wind
would be highly dependent of the tractor shape and the matching of the
tractor and trailer aerodynamics. The magnitude and unsteady nature of
these effects would vary greatly with the Reynolds number state
condition of the test. The importance of each discussed phenomena
would increase with advanced aerodynamic shaping.

5.5. Summary Findings


A review of capabilities and limitations of experimental test methods
employed to evaluate the aerodynamics of commercial vehicles
included wind tunnel test, coastdown, constant speed, and fuel
consumption testing. While wind tunnel testing is the only method
that obtains aerodynamic force and moment data all methods are
susceptible to Reynolds number effects and satisfying geometric and
flow similarity requirements. The importance of performing
Figure 5.12. Alignment of tractor/trailer in a non-banked curve. [228] aerodynamic testing at Reynolds number values in which the
boundary layer characteristics and aerodynamic forces are similar to
As seen in figure 5.12 the velocity vector and the resistance vector those of the full-scale vehicle at operational conditions were
are aligned with the tractor centerline that is yawed relative to the discussed along with the need to obtain measurement of the boundary
trailer. This effective yaw angle of the tractor and trailer will layer characteristics of the full-scale vehicle at operational condition.
introduce a nonlinear Reynolds number effect due to the variation in
yaw angle along the vehicle length that is not present at the same The importance of Reynolds number similarity to wind tunnel testing
frequency under operational conditions. This yaw angle sensitivity was discussed in the context of producing results representative of the
would be magnified by the allowed ambient wind conditions. operational vehicle. Recognizing that Reynolds similarity is complex
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 641

and dependent on many factors a series of criteria were provided for 276. Loving, D. and Katzoff, S., “The Fluorescent-Oil Film Method and
Other Techniques for Boundary-Layer Flow Visualization,” NASA
the comparison of wind tunnel to operational vehicles, operational MEMO 3-17-59L, 1959.
vehicle components and for comparing results between wind tunnel 277. Merzkirch, W., “Techniques of Flow Visualization,” AGARDograph No.
tests. To provide context an overview of Reynolds number 32, AGARD-AG-302, 1987.
sensitivities with wind tunnel facility type, floor/road simulation 278. “Blockage Corrections for Bluff Bodies in Confined Flows,”
Engineering Sciences Data Unit (ESDU) article 80024 November 1980
options, model-support considerations and with yaw testing.
279. “Wind Tunnel Wall Corrections,” AGARD-AG-336, Oct. 1998.
280. “Test Uncertainty,” ASME PTC 19.1-2005.
Road load testing options and guidelines were presented that indicate 281. “Experimental Validation and Uncertainty Analysis for Engineers,”
coastdown testing of commercial vehicles should require vehicle Coleman, H. and Steele, G., 3rd ed., Wiley, 2009
speeds to stay above 30mph and restrict wind speed to less than 5 282. ANSI/NCSL, Z540-2-1997, “U.S. Guide to the Expression of
Uncerainty in Measurement,” 1st ed., October 1997
mph. Underlying concern were outlined relative to Reynolds number 283. ISO 5725-1, “Accuracy (trueness and precision) of Measurement
state and the importance to tailor the coastdown and constant speed Methods and Results - Part 1: General Principles_and Definitions,”
test conditions to fall within the criteria for the specific Reynolds 284. Final Rulemaking to Establish Greenhouse Gas Emissions Standards and
Fuel Efficiency Standards for Medium- and Heavy-Duty Engines and
number state. This requires that the Reynolds number state of the test Vehicles, EPA-420-R-11-901, August 2011
vehicle at operational conditions of interest shall be determined prior 285. SAE International Surface Vehicle Information Report, “Aerodynamic
to testing. The following guidance is offered based on the vehicle Testing of Road Vehicles - Testing Methods and Procedures,” SAE
Standard J2084, Issued 1993.
operational conditions of interest; 1) for subcritical Reynolds number
286. The Oregonian published it online at http://www.oregonlive.com/
range operations the test shall not be constrained by Reynolds portland/index.ssf/2008/12/wind_tunnel.html, December 25, 2008
number 2) for transitional Reynolds number range operations a test 287. Leuschen, J. and Cooper, K., “Full-Scale Wind Tunnel Tests of
shall not be performed, and 3) for transcritical Reynolds number Production and Prototype, Second-Generation Aerodynamic Drag-
Reducing Devices for Tractor-Trailers,” SAE Technical Paper 2006-01-
range operations the test minimum width-based Reynolds number 3456, 2006, doi:10.4271/2006-01-3456.
shall be greater than 3 million. 288. Englar, R., “Improved Pneumatic Aerodynamics for Drag Reduction,
Fuel Economy, Safety and Stability Increase for Heavy Vehicles,” SAE
Technical Paper 2005-01-3627, 2005, doi:10.4271/2005-01-3627.
Overarching Reynolds number sensitivity criteria related to fuel 289. SAE International Surface Vehicle Information Report, “Cooling Flow
economy or fuel use testing was identified; 1) test conditions match Measurement Techniques,” SAE Standard J2082, Issued 1992.
the operational conditions of interest and 2) the vehicles at 290. SAE International Surface Vehicle Recommended Practice,
“Recommended Best Practice for Climatic Wind Tunnel Correlation,”
operational Reynolds number state is known prior to the test. The SAE Standard J2777, Issued 2007.
second criteria was highlighted as critical when interpreting test 291. Wiedemann, J. and Potthoff, J., “The New 5-Belt Road Simulation
results that may exhibit large run-to-run and segment-to-segment System of the IVK Wind Tunnels - Design and First Results,” SAE
Technical Paper 2003-01-0429, 2003, doi:10.4271/2003-01-0429.
variability resulting from an inability to maintain Reynolds number
292. Borello, G., Beccio, S., Gollo, G., and Quagliotti, F., “Rolling Road
similarity during the test. A specific area of concern is the use of oval Technology for Automotive High Speed Testing,” SAE Technical Paper
or circular test tracks. Data was presented that showed track testing 2000-01-0353, 2000, doi:10.4271/2000-01-0353.
can introduce an effective change (i.e. misalignment) in vehicle 293. Cogotti, A., “The New Moving Ground System of the Pininfarina Wind
Tunnel,” SAE Technical Paper 2007-01-1044, 2007, doi:10.4271/2007-
geometry that is not corrected for in either fuel use test method. A 01-1044.
discussion of the influence of subcritical and transitional Reynolds 294. Mau-Kuo, C., “Use of a Rolling Road System in Crosswind Conditions,”
number state on test data was discussed as well as the influence of Thesis, Old Dominion University, Pub No. 3574574, 2013
wind allowances in the J1321 and J1526 test protocols.5.6. 295. Hetherington, B., “Interference of Supports Used from Ground Vehicle
Wind Tunnel testing,” Thesis, Durham University, 2006
296. Hetherington, B. and Sims-Williams, D., “Wind Tunnel Model Support
Strut Interference,” SAE Technical Paper 2004-01-0806, 2004,
5.6. References doi:10.4271/2004-01-0806.
269. SAE International Surface Vehicle Recommended Practice, “Road 297. Simpson RL (2001), “Junction Flows,” Annual Review of Fluid
Load Measurement and Dynamometer Simulation Using Coastdown Mechanics, 33:415-443, January 2001.
Techniques,” SAE Standard J1263, Rev. 2010. 298. Shizawa, T., Honami, S., and Yamamoto, M., “Experimental Study of
270. SAE International Surface Vehicle Recommended Practice, “Road Load Horseshoe Vortex at Wing/Body Junction with Attack Angle by Triple
Measurement Using Onboard Anemometry and Coastdown Techniques,” Hot-Wire,” AIAA-96-0323. 34111 American Institute of Aeronautics &
SAE Standard J2263, Rev. 2008. Astronautics Aerospace Science Meeting, Reno, 1996.
271. Sandberg, T., “Heavy Truck Modeling for Fuel Consumption 299. Lombardi G., and Carmassi S., “Wall Interference Effects: Analysis
Simulations and Modeling,” Thesis No. 924, Linköping University, and Correction for Automotive Wind Tunnels,” MIRA Vehicle
2001. Aerodynamics, pp 1, Vol. 1, 2002
272. Fontaras, G., Dilara, P., Berner, M., Volkers, T. et al., “An Experimental 300. Wiedeman, J. and Ewald, B., “Turbulence Manipulation to Increase
Methodology for Measuring of Aerodynamic Resistances of Heavy Effective Reynolds Number in Vehicle Aerodynamics,” AIAA Journal,
Duty Vehicles in the Framework of European CO2 Emissions Vol. 27, No. 6, pp 763, 1989.
Monitoring Scheme,” SAE Int. J. Commer. Veh. 7(1):102-110, 2014, 301. “Blockage Corrections for Bluff Bodies in Confined Flows” Engineering
doi:10.4271/2014-01-0595. Sciences Data Unit (ESDU) article 80024 November 1980
273. Golsch, K., Duncan, B., and Kandasamy, S., “Improved CFD 302. Slangen, R., “Experimental Investigation of Artificial Boundary Layer
Methodology for Class 8 Tractor-Trailer Coastdown Correlation,” SAE Transition,” Thesis, Delft University of Technology, 2009.
Technical Paper 2013-01-2412, 2013, doi:10.4271/2013-01-2412.
303. Raghavan, G., “1D Transient Simulation of Heavy Duty Truck Cooling
274. Carr, G. and Stapleford, W., “Blockage Effects in Automotive System - HDEP 16 DST, Euro 6,” Thesis, Chalmers University of
Wind-Tunnel Testing,” SAE Technical Paper 860093, 1986, Technology, 2012.
doi:10.4271/860093.
304. Ng, E., “Vehicle Engine Cooling Systems: Assessment and Improvement
275. David, F. and Nolle, H., “Experimental Modeling in Engineering,” of Wind- Tunnel Based Evaluation Methods,” Thesis, RMIT University,
Butterworths, 1982 2012
642 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

305. Stephan, R., “Heat transfer measurements and Optimization Studies 314. Walston, W., Buckley, F., and Marks, C., “Test Procedures for
Relevant to Louvered Fin Compact Heat Exchanges,” Thesis, Virginia the Evaluation of Aerodynamic Drag on Full-Scale Vehicles in
Polytechnic and State University. Windy Environments,” SAE Technical Paper 760106, 1976,
306. Wickern, G., Zwicker, K., and Pfadenhauer, M., “Rotating Wheels - doi:10.4271/760106.
Their Impact on Wind Tunnel Test Techniques and on Vehicle Drag 315. Lucas, G. and Britton, J., “Drag Data from Deceleration Tests and Speed
Results,” SAE Technical Paper 970133, 1997, doi:10.4271/970133. Measurement During Vehicle Testing,” University of Loughborough,
307. Elham, A., Van Raemdonck, G. and van Tooren, M., “Design, 1970.
construction, and validation of a new boundary layer rake for full-scale 316. White, R. and Korst, H., “The Determination of Vehicle Drag
testing,” Journal of Flow Measurement and Instrumentation, 26 (2012) Contributions from Coast-Down Tests,” SAE Technical Paper 720099,
30-36 1972, doi:10.4271/720099.
308. Bui, T., Oates, D. and Gonzalez, J., “Design and Evaluation of a New 317. Roussillon, G., Marzin, J. and Bourhis, J., “Contribution to the Accurate
Boundary-Layer Rake for Flight Testing,” NASA TM-2000-209014, Measurement of Aerodynamic Drag by the Deceleration Method,” Br.
2000. Hydrodyn. Res. Assoc., Adv. Road Vehicle Aerodyn. Pap. 4, 1975.
309. HS-1566, “Aerodynamic Flow Visualization Techniques and 318. Passmore, M. and Le Good, G., “A Detailed Drag Study Using
Procedures,” (Warrendale, Society of Automotive Engineers, Inc., 1986), the Coastdown Method,” SAE Technical Paper 940420, 1994,
ISBN 978-0-89883-418-5. doi:10.4271/940420.
310. Hoerner, S., “The Determination of Aerodynamic Resistance of Vehicles 319. Buckley, F., “ABCD - An Improved Coast Down Test and Analysis
from Free Motion Method,” VDI Z., 1935, 79, 1028-1033. Method,” SAE Technical Paper 950626, 1995, doi:10.4271/950626.
311. Roussillion, G., “Contribution to Accurate Measurement of 320. Saltzman, E. and Meyer, R., “Drag Reduction Obtained by Rounding
Aerodynamic Drag on a Moving Vehicle from Coast-Down Tests and Vertical Corners ion a Box-Shaped Ground Vehicle,” NASA TM-X-
Determination of Actual Rolling Resistance,” Journal of Wind Engr. And 56023, 1974.
Ind. Aerodynamics, 9 (1981) 33-48. 321. Walter, J., Pruess, D., and Romberg, G., “Coastdown/Wind Tunnel Drag
312. Roussillon, G., Marzin, J. and Bourhis, J., “Contribution to the Accurate Correlation and Uncertainty Analysis,” SAE Technical Paper 2001-01-
Measurement of Aerodynamic Drag by the Deceleration Method,” 0630, 2001, doi:10.4271/2001-01-0630.
Advances in Road Vehicle Aerodynamics 1973, Stephens H. S., ed., 322. 40 CFR Part §1066 - Vehicle-Testing Procedures, Subpart D-Coastdown
1973, pp. 53-62. Procedures for Heavy-Duty Vehicles (1066.301 - 1066.315)
313. Lucas, G. and Emtage, A., “A new Look at the Analysis of Coast-Down
Test Results,” Proceedings of the Institution of Mechanical Engineers,
Part D: Journal of Automobile Engineering 1987 201: 91, DOI:10.1243/
PIME_PROC_1987_201_163_02
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 643

CHAPTER 6. VEHICLE AERODYNAMIC having to respond to many national and international economic
DESIGN events as well as revolutionary technological advancements. This
success is attributed to the evolutionary and at times revolutionary
The initial pursuit of low-drag commercial vehicle designs were based
advancement in all disciplines and technology sectors that comprise a
on streamlining techniques and first appeared in the 1930's [90, 105,
commercial ground vehicle, of which aerodynamics in one. Over the
106, 107] and continued through the 1940's, for example see Labatt
past four decades the aerodynamic drag coefficient of a heavy
Brewing Company vehicle of 1947 [127]. Even though these
combination vehicle has been reduced by greater than fifty percent
aerodynamic demonstrations did not impact the shape of the typical
through the use of established drag reduction technology guided by
in-service vehicle they did serve as the seed corn for the efforts that
the application of both computational and experimental methods in
followed. Leveraging findings from the streamlining studies were
the design process. A graphical depiction of these trends is depicted
several notable wind tunnel experiments from 1950 through the 1970's.
in figures 6.2, 6.3, and 6.4 [2, 4, 5, 40, 41, 77, 127, 133,160, 170,
These test programs worked with representative contemporary vehicle
183, 246, 323, 324, 325, 326, 327].
concepts and used a systematic approach to develop aerodynamic
concepts and technology databases. Most relevant of these are the work
Figure 6.2 presents the drag coefficient trend from 1970 to 2030 for a
by Trailmobile and University of Maryland in the 1950's [125, 126],
commercial vehicle and for reference purposes an automobile [324].
Flynn and Kyropoulos in 1960's [32] and Mason and Beebe in 1970's
These curves were constructed to reflect the general characteristics of
[323]. Each of these investigations were performed at a width based
the majority of operational vehicles at the time and may not represent
Reynolds number greater than 1.0 million and their results
the full range of in-service vehicle drag coefficients. Market forces
demonstrated significant drag reductions with simple aerodynamic
drive the lower boundary for each trend and the lagging upper
fairings, as shown in figure 6.1. Each of these studies demonstrated a
boundary is a result of regulatory criteria. The decrease in commercial
fifty percent drag reduction from the baseline vehicle. The 1970 paper
vehicle drag coefficient between 1970 and 2000 is a result of reduced
by Ludvigsen [119] and Cooper's 2004 paper [127] provide additional
flow separation attributed to tractor shaping and fairings in combination
details on the early history of commercial vehicle aerodynamic design.
with minor changes in trailer geometry [1, 2, 19, 20, 21]. The decrease
from 2000 to 2015 results from a combination of additional tractor
aerodynamic shaping and the addition of aerodynamic fairings to the
tractor and trailer [2, 4, 5, 52, 65, 66]. For reference, estimated lower
drag limits for a practical commercial vehicle [134] and automobile
[324] are placed on the figure. The increasing width of the commercial
vehicle trend for 2015 and beyond reflects the diversity in vehicle type
demanded by the industry [327]. The lower border of the commercial
vehicle trend band is representative of demonstration vehicles that may
become production models in response to regulatory actions [67, 284].
A comparison of the two trend bands show that the commercial
vehicles sector is at least 15 years behind the automotive sector in
achieving a equivalent level of drag reduction relative to the estimated
lower limit. The ability to achieve the additional drag reduction beyond
2015 will be dependent on the management of both pressure and
Figure 6.1. Photographs of tractor-trailer model tested in the University of friction drag forces as noted in Figure 6.3.
Maryland wind tunnel. [125,126]. Photographs obtained from Cooper [127].
Figure 6.3 presents commercial vehicle trend lines for the pressure
It is worth noting that it was not until Flynn's 1962 paper [32] that a
and friction drag components that comprise the drag coefficient trend
discussion of Reynolds number effects weas published. The importance
line shown in figure 6.2. The pre-2000 drag split is approximated by a
of Reynolds number in ground vehicle aerodynamics was also raised by
flat-plate based analysis of the wetted area for a typical combination
those in attendance at the General Motors sponsored Symposium entitled
vehicle at 60mph in concert with published data [328]. The post-2000
“Aerodynamic Drag Mechanisms of Bluff Bodies and Road Vehicles”
increase in friction drag results from the decrease in total drag and the
held in 1976 [54]. It is unfortunate that these early Reynolds number
increase in total wetted surface area resulting from the addition of
conversations did not stimulate detailed investigations of Reynolds
aerodynamic fairings discussed previously. These trend lines clearly
number sensitivity for this class of vehicle. While it is generally agreed
highlight the growing significance friction drag will play for current
that data from these early studies was critical in demonstrating the
and future vehicle designs. The chart indicates that further
potential of aerodynamic treatments to reduce pressure drag and thereby
improvements in aerodynamic drag will require reductions in both
fuel consumption the experimental set-up, instrumentation and baseline
friction as well as pressure drag requiring Reynolds number similarity
vehicle geometry did not focus their effort to satisfy flow similarity
criteria to be rigorously satisfied in both experimental and
criteria required for the assessment of Reynolds number sensitivities and
computational investigations. The availability and capability of the
little attention was given to assessing friction drag contributions.
experimental and computational tools will determine the success of
future aerodynamic studies. Unfortunately, a literature review of
Since 1970 the commercial vehicle industry has consistently
commercial vehicle aerodynamics indicates a shifting imbalance
developed vehicles that meet and in most cases exceed the needs of
between experimental and computational efforts is underway as
its customers and stringent regulatory requirements [67, 284] while
depicted in figure 6.4.
644 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

complete vehicle. This trend is also supported by the availability


and cost of this data source. However, there is concern that
computational tools may not be able to satisfy Reynolds number
similarity criteria without the development of commercial vehicle
specific boundary layer models. In contrast the experimental
projection is much more complex as noted by the expansion of the
trend curve of figure 6.4 after 2015. The upper boundary of the
curve could be realized if current testing procedures are adjusted to
satisfy Reynolds number similarity criteria. The lower boundary
reflects a continuation of current testing practices. To ensure quality
and continuity of both experimental and computational results the
industry is encouraged to routinely update SAE Recommended
Practices J2966 [133] and J1252 [64], develop a CFD verification
and validation standard [232, 233, 234] and launch a drag
prediction workshop [263].

Figure 6.2. Historical trend line for the variation in vehicle drag for ground
vehicles.

Figure 6.4. Historical variation in the publication of computational and


experimental aerodynamic data for commercial vehicles.

Failure by the commercial vehicle community to react to these


changing needs could result in an outcome faced by the aircraft
Figure 6.3. Historical trend line for the variation in pressure and friction drag community nearly 30 years ago when then Aerodynamics Chief at
for commercial vehicles. NASA Langley Research Center predicted [329]:

Shown in figure 6.4 are trend lines depicting the percentage of “I see in the year 2020, there will be no wind tunnels. I would say we
published commercial vehicle aerodynamic studies containing would be at the point where airplanes could be designed by rather
experimental and/or computational data. A primary source for these low-paid technicians.” Dr. Doug Dwoyer, 1987.
findings is the archives of NASA, DOE, SAE, AIAA, ASME and
academia as discussed in Chapter 1. The trend lines show the shift The following subsections of this chapter will expand on the
from an experimental dominated industry, prior to 1990, to one that discussions related to figures 6.2, 6.3 and 6.4 and review the evolution
relies more heavily on computational than experimental tools in of Reynolds number related findings and guidelines from 1970 to
2015. Projecting these trends beyond 2015 is challenging due to the 2015. A review of vehicle design trends will discuss the increased use
lack of clearly defined constraints, such as the GHG regulations [67], of aerodynamic shaping to increase vehicle performance. These trends
that impact the drag trends shown in figure 6.2 and 6.3. The post- highlight the importance of Reynolds number, boundary layer
2015 trends are constructed based on; the drag trends discussed management and flow separation control for current and future
previously, availability and cost of each data source and the ability of vehicles. The final subsection will present future Reynolds number
each data source to satisfy Reynolds number similarity criteria. challenges for the 2015 to 2030 time period.

Computational analysis is expected to be included in all future


vehicle investigations within the next five years as they are the only
means to differentiate between friction and pressure drag for a
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 645

6.1. Laying the Foundation: 1970 to 2000 facilities [5] are shown in figure 6.6 [127]. A review of the NRC
The three decades from 1970 to 2000 saw numerous national and publications suggest that the wind tunnel test program focused on
international events that served as an impetus for the evolution in three topics; 1) evaluation of wind tunnel testing methods, 2)
commercial vehicle aerodynamics, see Section 1.4 of Chapter 1. Most evaluation of aerodynamic fairings to maximize drag reduction, and 3)
notable were the 1973 oil embargo and the 1979 energy crisis that comparison of wind tunnel data with full-scale coastdown and fuel
stimulated numerous aerodynamic and fuel economy improvement consumption test findings. Results from the NRC drag reduction
studies by NASA, EPA, DOT, ATA, SAE and the NRC in Canada [1, studies (see left side of figure 6.6) showed a fifty percent reduction in
23, 24, 25, 26, 45, 53, 88, 205, 206, 207, 208, 320]. Participating in drag compared to the baseline vehicle resulting in a drag coefficient
these efforts were most commercial vehicle manufacturers as well as for the fully treated vehicle between 0.40 and 0.50. These results are
numerous Universities [18, 19, 20, 21, 35, 41, 50, 57, 58, 189, 319]. consistent with those from Trailmobile and University of Maryland
[125, 126], Flynn and Kyropoulos [32] and Mason and Beebe [323].
The 1970-era vehicles were sharp-edge bluff bodies with
aerodynamic characteristics dominated by large regions of separated
flow resulting in high levels of pressure drag. Based on historical data
this class of vehicle exhibits minimal sensitivity to changes in
Reynolds number and boundary layer state, as measured by changes
in drag coefficient [2, 3, 32, 38, 40].

In 1973 NASA stepped into the fray and focused their efforts on the
fundamental aerodynamics of bluff body ground vehicles [88]. The
NASA effort recognized the importance of Reynolds number and
structured a program that focused on full-scale vehicles (see figure 6.5)
and employed the coast-down testing method of Hoerner [38]. This
strategy eliminated Reynolds number concerns and through detailed
systematic testing they were able to investigate the drag reduction
potential of current model vehicles as well as optimum shaped bluff
body ground vehicles. Testing investigated tractor shaping, gap effects, Figure 6.6. Photographs of NRC wind tunnel test articles [126].
edge rounding and boattails, see publications by NASA [48, 53, 205,
206, 207, 208, 209, 320], AIAA [45], and SAE [186]. This testing and A significant contribution during this time period was the 1979
subsequent analysis contributed to defining the lower drag limit for publication of SAE Recommended Practice J1252 entitled “Wind
combination vehicles as depicted in figure 6.2 [88]. Tunnel Test Procedure for Trucks and Buses” [62]. This document
provided the first set of wind tunnel testing criteria for commercial
vehicles and established the original set of Reynolds number
guidelines. Surprisingly, both the original 1979 version and the 1981
update of the document recommend a minimum test Reynolds
number value of 0.7 million, see horizontal green dashed line in
figure 6.7. Figure 6.7 depicts the historical trend in industry
recommended minimum Reynolds number for wind tunnel testing.
The figure shows industry recommendations based on published data
(circle symbols), values contained in SAE Standard J1252 (green
symbols and dashed lines) and values set by U.S. GHG regulations
(orange symbols and dashed lines. Note, the 0.7 minimum test
Reynolds number was not supported by references in the J1252
document and the value less than the test Reynolds number used for
the majority of vehicle testing efforts at the time. Immediately
following the release of the 1981 revision of the J1252 document a
Figure 6.5. Photographs of NASA test vehicles [88]. number of leading aerodynamicist raised concerns with the 0.7
million value. In 1981 Gilhaus proposed 1.6 million [35], Cooper
In parallel with the fundamental NASA studies a large number of wind proposed 2.0 million in 1982 [23] and 1984 [330], Drollinger
tunnel test of conventional commercial vehicles were being performed suggested 2.0 million in 1987 [1], Olson offered 1.7 million in 1992
in the U.S, Canada and Europe [24, 25]. Aerodynamic studies at NRC [46] and in 1998 Hucho suggested 2.5 million [40]. These values are
[23, 24, 25, 26, 27, 166, 167, 168, 202] are representative of the graphically depicted as SAE Journal data points in figure 6.7. Despite
diversity of wind tunnel testing performed at the time. Photographs of the response from industry the minimum wind tunnel test Reynolds
full-scale and sub-scale wind tunnel test articles installed in NRC number value of 0.7 million remained unchanged.
646 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

(1986 version) [336] fuel consumption test procedure. The EPA


program focused on combination vehicles and required track testing,
see figure 6.9. CARB realized they lacked the resources to
independently evaluate trailer aerodynamic technology and required
industry to use only aerodynamic products that had been accepted and
ranked by the EPA. This action by CARB transformed the “voluntary”
EPA Smartway program [332] into a legislated program and created a
significant increase in EPA Smartway testing as aerodynamic product
manufactures pursued the CARB created marketplace.

The industry has made significant aerodynamic improvements over


the previous three decades as reflected in figures 6.2, 6.3, and 6.4.
In response to rising fuel costs after 2000 and other market forces
the commercial vehicle industry began accelerating implementation
of aerodynamic technology. This evolution increased the
Figure 6.7. Historical trend in recommended minimum Reynolds number for
wind tunnel testing
importance of boundary layer and flow separation control in drag
reduction efforts. Significant Reynolds number related activities
By 1990 the flurry of aerodynamic studies, driven by the national 55 from 2000 to 2010 grew from the DOE activity that produced a
mph speed limit and the energy crisis, were winding down and number of significant contributions [4, 10, 15,17, 25, 27, 51, 52, 55,
aerodynamic design concepts had gained a foothold in the industry. 56, 127, 158, 159, 160, 169, 171, 173, 175, 177, 187, 200, 213, 214,
Truck cabs were undergoing gradual evolution to more aerodynamic 269]. Most notable was the Reynolds number investigation by
shaping, edge rounding and tractor roof fairings became the norm and Storms [55] as presented in figures 3.13 and 3.14. The analysis of
computational methods were being used in the design process. These Storms data suggest a minimum test Reynolds number of 3.0
modest aerodynamic improvements pushed drag coefficients below million, see 2004 data point in figure 6.10. European aerodynamic
0.70 by 2000, see figure 6.2. studies provided additional insight into the importance of Reynolds
number to heavy truck aerodynamics. In 2009 Hjelm [197]
presented a paper entitled “European Truck Aerodynamics - A
6.2. Recent Events: 2000 to 2015 Comparison Between Conventional and CoE Truck Aerodynamics
A surge of aerodynamic activity occurred in the 2000 to 2015 time and a Look into Future Trends and Possibilities” at the The
period brought about by; rising fuel prices, a significant influx of Aerodynamics of Heavy Vehicle II: Trucks, Buses and Trains
funding from the DOE 21st Century Truck Program in 1998 [63] and conference which showed a minimum test Reynolds number of 4.0
DOE SuperTruck Program in 2010 [331], fleet support from the million, see 2009 data point in figure 6.10. These values are
Smartway Program in 2004 [331], and regulatory requirements consistent with pre-2000 findings and contradict the 0.7 million
imposed by the CARB Tractor-Trailer GHG Regulation in 2010 [66] minimum test Reynolds number value of J1252 (see horizontal
and EPA/DOT GHG in 2011[285]. The overarching objective of these green dashed line in figure 6.10). Note, a 0.7 million Reynolds
programs was to reduce the energy use and GHG emissions of number corresponds to a vehicle operating at 8.5 mph. These
commercial vehicles. A discussion of the aerodynamic specific goals experimentally derived minimum Reynolds number values
can be found in references 331 indicated the need to revise SAE J1252 and suggest a reevaluation
of Reynolds number effects for commercial vehicles.
From 2000 to 2010 DOE and EPA goals were similar but their
approach to aerodynamics varied greatly. DOE pursued fundamental Beginning in 2010 the DOE began to transition from the 21st Century
wind tunnel, coastdown and computational simulation studies of bluff Truck program to a demonstration activity with the launch of the
body and commercial vehicle aerodynamics in partnership with SuperTruck program [331] and the EPA/DOT initiated the
NASA, NRC, academia and other noted experts in the industry, see development of the Phase I GHG regulations [333]. To support
figure 6.8 [177]. The EPA/DOT performed studies to guide pending GHG regulations the SAE launched a task force to update
modifications to SAE automobile coastdown procedures J1263 [274] J1321 fuel use test document [134] and recognizing that J1263 and
and J2263 [275] to support pending GHG regulations. The primary J2263 are fundamentally flawed for testing heavy trucks a new task
use of the coastdown procedures by EPA was the extraction of an force was launched to develop a commercial vehicle coastdown test
aerodynamic drag coefficient for an operational vehicle. procedure (J2978). The J2978 document remains in development at
this time with expected publication in 2016. The revised J1321 was
In contrast EPA Smartway managed a voluntary test program that published in 2012. The revisions clarified testing criteria for
accepted/rejected and ranked trailer aerodynamic fairings. The EPA aerodynamic technology, defined aerodynamic influences on test data,
program required all testing to use the EPA modified SAE J1321 and addressed data quality concerns.
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 647

Figure 6.8. Photograph of 0.25 scale heavy truck model installed in the NASA Langley Research Center 30×60 foot wind tunnel [177].

Figure 6.9. Photograph of heavy truck undergoing J1321 test at the Transportation Research Center.

published in 2013. Both J1252 and J2966 provide specific guidance


on Reynolds number simulation with a focus on design and
development of operational vehicle. Based on published data the 0.7
million Reynolds number was removed from J1252 and replaced
with a detailed test procedure to determine the minimum allowable
Reynolds number for the specific test being conducted. The intent
of the test Reynolds number assessment procedure is to allow
adjustments to the minimum test Reynolds number based on
differences in the test vehicle or the specific operational Reynolds
number range (see Chapter 3, section 3.1) of the reference full-scale
vehicle. The 2012 version of J1252 is depicted in figure 6.10 by the
vertical green dashed line spanning the shaded region for year 2012.
The vertical extent of this line reflects the range of feasible
minimum Reynolds number values for commercial vehicles. The
J1252 recommended practice also provides guidance in the
Figure 6.10. Historical trend in recommended minimum Reynolds number for execution and interpretation of the minimum allowable Reynolds
wind tunnel testing number procedure results as well as criteria and test corrections for
test model boundary layer as well as wind tunnel floor and wall
Recognizing the growing importance of aerodynamics to the boundary layers.
industry, SAE formed task forces in 2009 to revise the J1252 wind
tunnel test document [64] and in 2011 to develop a new document, Over this same time period several studies were published that
SAE J2966, Guidelines for Aerodynamic Assessment of Medium further clarified both the test Reynolds number topic and Reynolds
and Heavy Commercial Ground Vehicles Using Computational number sensitivity effects, as noted by the three grey and one white
Fluid Dynamics [133]. J1252 was published in 2012 and J2966 was circle symbols for years 2012 to 2013 in figure 6.10. In 2012 Wood
648 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

[2] suggested 3.0 million based on a review of published data from vehicle, see figures 6.12 and 6.13, and the 2011 paper investigated a
1950 to 2010 and three significant wind tunnel test reports were similar limit for a 2010 model single unit straight truck, see figure
published in 2013. The SAE paper by McArthur [66] recommends a 6.14, 5.15 and 6.16.
value greater than 2.4 million and Leuschen's SAE paper [65]
recommends 4.0 million. Leuschen's recommendation is based on The combination vehicle test employed a 0.25 scale model of a
test results for full-scale operational vehicles and 0.50 scale models representative tractor-trailer, see photographs shown in figure 6.12.
of the same vehicle in the NRC 9-meter wind tunnel [5] over a Rew Presented in figure 6.13 is a plot of the variation in drag with yaw
range of 1.0 to 6.25 million. McArthur's study was performed in the angle for each aerodynamic device installed on the test model.
Monash wind tunnel and investigated a 0.33 scale model over a Rew Facility wind speed restriction limited the test Reynolds number for
range of 0.27 to 2.4 million. the 0.25 scale model to 1.26 million. As a result a Reynolds number
sweep was performed to determine if the drag had reached a constant
The final paper of note was a DOE sponsored test in the NASA Ames value at 1.26 million. In parallel with the Reynolds number
Research Center (ARC) 80 × 120 foot atmospheric wind tunnel [181] sensitivity assessment the boundary layer on the trailer was evaluated
in which an full-scale vehicle was tested at wind speeds from 20mph with a boundary layer survey rake. This assessment indicated that the
(Re, w = 1.6 million) to 80 mph (Re, w = 6.4 million) with a nominal minimum test Reynolds number criteria was not satisfied and the
test speed of 58 mph (Re, w = 4.6 million). Shown in figure 6.11 is a model boundary layer had not become fully turbulent at the model
photograph of the vehicle installed in the NASA wind tunnel and a trailing edge. The lack of a fully turbulent boundary layer at the
Reynolds number against drag sensitivity plot for the baseline vehicle. trailing edge would impact the relevance of the base drag
The Reynolds number data plot was taken directly from the subject measurement and boattail performance. The test program addressed
report and show drag results for a limited set of tested Reynolds these concerns through the use of artificial roughness, as noted by the
numbers due to proprietary concerns by DOE. The plot show the four vertical grey strips on the trailer. Roughness elements were
minimum Reynolds number is significantly greater than 2.0 but less positioned on the trailer sides and top to force transition in order to
than 4.6. This is noted in figure 6.10 by the white circle for 2013. match the full-size turbulent boundary layer character at the trailer
base and minimize Reynolds number sensitivity. Boundary layer
surveys of the trailer boundary layer confirmed the test objectives
were met. Based on the trailer boundary layer study and prior to the
test start artificial roughness was also applied to the tractor. A review
of the drag data presented in figure 6.13 show results that are
consistent with those obtained by NRC [23, 24, 25, 26, 27, 166, 167,
168, 202] and DOE [181].

b. Plot of the change in wind averaged drag with increasing Reynolds number.

Figure 6.12. Quarter scale generic combination vehicle installed in the


LFST. [177]

In contrast to the combination vehicle test the single unit straight


truck test was performed at full-scale Reynolds numbers with an
operational vehicle [176]. For this study the average operational
a. Photograph of full-scale vehicle installed in the wind tunnel. speed was assumed to be 45 mph resulting in a Reynolds number of
3.47 million. The vehicle used for testing was a GMC T-series
Figure 6.11. Reynolds number sensitivity information for a full-scale vehicle
installed in the NASA ARC 80 × 120 foot wind tunnel. [181] straight truck as shown in Figure 6.14. The truck features a modern
cab-forward design and a standard 16-foot box with radiused corners
Dr. Landman authored two additional relevant publications in 2010 and edges. Additionally, it should be noted that the vehicle includes a
[177] and 2011 [176]. Both papers reported wind tunnel test results lift-gate at the rear of the vehicle. Several devices were chosen for
obtained in the Langley Full Scale Tunnel (LFST). The 2010 paper this test including a front valance, a bubble fairing for the front of the
investigated the drag reduction limit for a representation combination box, a boat-tail at the rear of the box, an ideal side skirt, and a
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 649

practical side skirt. The devices were tested in a “buildup” degree yaw, and 10 percent increase in CD at 10 degree yaw.
methodology where treatments were progressively added to the Comparing these results with those for a combination vehicle
vehicle. Highlighted in red on the photographs are the various vehicle depicted in figure 3.14 show similar magnitude in the change in drag
treatments. Listed in the legend of the data plot are the various but an opposite trend. Where the combination vehicle showed a
aerodynamic treatments. The maximum drag reduction achieved is decrease in drag with increasing Reynolds number the single unit
equivalent to that obtained for combination vehicles and consistent truck shows an increase in drag. More significant is the observation
with that reported in references 169 and 337 for straight trucks. that the drag of the single unit truck does not approach an asymptotic
state over the Reynolds number range. These results show that this
vehicle is operating in a transitional Reynolds number range at speeds
between 16 and 80 mph and never achieves a transcritical Reynolds
number condition (see Chapter 3). The data of figures 6.14 to 6.15
show that wind tunnel testing of this vehicle type must be performed
at full-scale Reynolds numbers.

The final 2010 to 2015 events to discuss are the implementation of


GHG regulations, represented by the orange symbols and dashed line
in figure 6.10. Despite the extensive body of work and the availability
of accepted industry standards addressing Reynolds number
sensitivity the U.S. regulatory bodies have published Reynolds
number guidelines that are not supported by published data. In 2012
CARB GHG program published a wind tunnel test procedure [67]
Figure 6.13. Drag coefficient data for various aerodynamic treatments on the that set a minimum test Reynolds number of 1.0 million and requires
¼ scale generic combination vehicle. [177] test models to be 0.125 scale or larger, see 2012 orange symbol. The
restriction of model scale to 0.125 and larger is surprising as a
significant body of the trailer aerodynamic fairing developed was
performed at NRC on 0.10 scale models [23, 24, 25, 26]. Neither of
these criteria is supported by published references and they directly
contradict the guidance provided in the 2012 version of J1252.
Following CARB's lead the EPA Smartway program began accepting
aerodynamic data based on CARB's criteria. In 2015 the EPA and
NHTSA Propose Standards to Reduce Greenhouse Gas Emissions
and Improve Fuel Efficiency of Medium- and Heavy-Duty Vehicles
for Model Year 2018 and Beyond also adopted the CARB criteria, see
orange symbol and horizontal dashed line at year 2015. The GHG
Reynolds number criteria and test guidelines will introduce
significant errors in the assessment of aerodynamic drag that are an
order of magnitude greater than the data uncertainty requirements
imposed by these programs.

Figure 6.14. Photographs and drag coefficient data for a full-scale single unit
truck. [176]

The testing of a full-size vehicle allowed for an investigation of


Reynolds number effects as shown in figures 6.15 and 6.16. Shown in
figure 6.15 is the variation in drag with yaw angle for Reynolds
numbers of 1.29, 3.47 and 4.93 million representing vehicle speed of
16.7, 45.0 and 65.0 mph. These data show a significant increase in
drag coefficient with increasing Reynolds number at all yaw angles.
The shaded regions on the figure highlight the increasing increment
in CD due to increasing Rew for yaw angle of 0, 6 and 12 degrees. To
investigate this observed trend the data of figure 6.15 is plotted as the Figure 6.15. Effect of Reynolds number on Yaw dependent CD for a full-scale
fractional change in DCD, with Rew for three yaw angles. The figure single unit truck. [176]
shows that increasing Rew from 1.29 to 4.93 million results in a 7
percent increase in CD at 0 degree yaw, 8 percent increase in CD at 6
650 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

[135] and develop both Coastdown and Constant speed standards


specifically designed to address the unique test needs of commercial
vehicles. However, the actions of the regulatory community highlight
the need for additional refinements to the SAE commercial vehicle
standards to clarify the importance and impact of Reynolds number
effects on aerodynamic and fuel consumption characteristics.

6.3. Future Challenges: 2015 to 2030


Fuel costs and GHG regulations continue to place a heavy burden on
the freight industry. Numerous studies and demonstration projects
have shown that the application of aerodynamic technology can
provide significant improvement in fuel economy [181].
Improvements in vehicle aerodynamics and drag reduction may
introduce unintended consequences and opportunities related to
vehicle design and systems integration. Areas of opportunity are;
engine downsizing, improved engine and brake cooling flow, reduced
Figure 6.16. Variation in the change in CD with Reynolds number for a
splash and spray, enhanced aerodynamic stability and increase traffic
full-scale single unit truck. [176]
safety due to reduced interference flows. In contrast, unintended
The historical trends of figure 6.10, the data presented in figures 6.2, consequences include; brake system upsizing, increased maintenance,
6.3, 6.4 and 6.11, 6.12, 6.13, 6.14, 6.15, 6.16, and published findings and large variability in aerodynamic resistance due to atmospheric
[1, 2, 3, 4, 5, 10, 15,17, 18, 19, 20, 21, 23, 24, 25, 26, 27, 32, 35, 38, interference effects. The ability to maximize the opportunities and
40, 41, 45, 46, 48, 50, 51, 52, 53, 54, 55, 56, 57, 58, 65, 66, 88, 119, eliminating the negative possibilities requires a thorough
125, 126, 127, 133, 158, 159, 160, 166, 167, 168, 169, 171, 173, 175, aerodynamic evaluation of the vehicle and its aerodynamic dependent
176, 177, 181, 186, 187, 189, 197, 200, 202, 205, 206, 207, 208, 209, systems under operational conditions.
213, 214, 269, 319, 320, 323, 324, 325, 326, 327, 328, 329, 330, 337]
show that test Reynolds number impacts both data trends and Referring to the historical trends of figure 6.17 (combination of
magnitudes. It has been shown that test or analysis results obtained at figures 6.2, 6.3 and 6.4), the majority of available fuel savings from
less than full-scale Reynolds numbers will differ from drag levels at pressure drag reduction has been realized through the extensive use
full-scale Reynolds number by more than ten percent. The impact of of traditional fairings, see figure 6.17a [23, 27, 52, 160, 176, 177,
Reynolds number on vehicle aerodynamics is very complex and will 181, 338, 339, 340, 341, 342]. Published data suggest that the
vary with vehicle shape, atmospheric conditions, aerodynamic industry average drag coefficient of next generation in-service
treatments, and both vehicle speed and yaw angle. To address these combination vehicles will be less than 0.50 with a potential to
critical factors the SAE has revised J1252 [64] and J1321 [134] and achieve 0.40 by 2030. One byproduct of the pressure drag reduction
created J2966 [133]. The SAE has activity underway to revise J1526 is seen in figure 6.17b where the relative importance of friction drag
will increase as the pressure drag is reduced.

Figure 6.17. Historical drag trends for combination vehicles.


Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 651

designs. Of particular importance is the maturation of computational


simulation tools with improved friction modeling capability
specifically developed for commercial ground vehicles. The following
subsections will discuss each of the three drag reduction topics.

Figure 6.18. Historical trend of pressure and friction drag for a combination
vehicle.
Figure 6.19. Variation in crossover speed with drag coefficient and vehicle
As shown in figure 6.18 the fairings used to reduce pressure drag
weight for a combination vehicle.
increases vehicle-wetted area resulting in an increase in absolute
friction drag level, see figure 6.18. Noted on the figure is the trend line
from 1970 to 2015 showing the increase in wetted area attributed to 6.3.1. Friction Drag
aerodynamic treatments added to a combination vehicle, compared to A defining difference between a streamline/aerodynamic body and
the wetted area of a baseline vehicle without aerodynamic treatments. bluff-body is the ratio of pressure drag to friction drag. If the ratio of
pressure to friction drag is greater than 1 we define the body as bluff
The graphic show a minimal increase in wetted area, compared to the and if the ratio is less than 1 the body is streamline/aerodynamic.
baseline vehicle of the time, between 1970 and 2000 followed by a Historically, commercial vehicles have been clearly defined as a
twenty percent increase in wetted area from 2000 to 2015. This bluff-body however the evolution of this vehicle class is clearly
increase is primarily attributed to the extensive use of trailer fairings changing this perspective. If the current trend continues commercial
and to a lesser degree to tractor fairings. vehicles will approach a drag ratio of 1 and may be defined as a
streamline bluff-body (i.e. an aerodynamic oxymoron).
A second byproduct of a reduced drag coefficient is a decrease in fuel
savings for given percent reduction in vehicle drag. The current rule Friction drag reduction concepts can be used to reduce the turbulent
of thumb in the industry is a two percent reduction in drag coefficient friction coefficient and promote increased laminar flow by
will produce a one percent fuel savings. The speed at which the 2:1 suppressing boundary layer transition. Friction drag may also be
ratio is valid is termed the crossover speed. Where crossover speed is reduced in a synergistic design process where flow control systems
defined as the vehicle speed at which the aerodynamic drag force is are used to improve fairing performance allowing for the
equal to a combination of rolling resistance and driveline losses. minimization of fairing surface area that further reduce friction drag.
Pursuing these advanced flow management concepts requires an
At speeds below the crossover speed the fuel savings falls below the understanding of the boundary layer flows present on a commercial
2:1 ratio. Figure 6.19 shows trend lines for crossover speed with drag vehicle and any variations in friction drag and limitations associated
coefficient for three vehicle weights. The chart shows a 60,000 pound with sub-scale wind tunnel testing.
vehicle, with a 0.60 drag coefficient, has a 53 mph crossover speed
and reducing the drag coefficient to 0.40 will increase the crossover A flat plate based analysis of the friction drag characteristics on a
speed to 65 mph. Reducing the vehicle drag coefficient to the combination vehicle is presented in figures 6.20 and 6.21. Figure 6.20
conceptual lower limit (vertical green line at CD of 0.27) would push is a plot of the average friction coefficient and figure 6.21 shows the
the crossover speed to 80 mph. variation in the average friction drag coefficient for a full-scale
vehicle and 0.250 scale and 0.125 scale models for Reynolds number
To achieve drag reduction beyond 2015 will require an all-inclusive from 0.5 million to the maximum test Reynolds number based on the
vehicle system level design perspective rather than summing compressibility limit.
individual technology contributions. Aerodynamic drag reduction will
require the use of friction drag reduction concepts, flow control The calculation of the friction coefficient presented in figure 6.20
systems, and management of atmospheric and vehicle interference assumes a smooth surface and a transition Reynolds number of 1.0
flows. Critical to the success of each topic area is the continued million. The transition Reynolds number value is used to define the
development of aerodynamic test and analysis tools capable of laminar and turbulent portion of the wetted area. A review of the
capturing the Reynolds number sensitivities present with future friction coefficient curves show that the full-scale vehicle boundary
652 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

layer would be predominately turbulent for all Reynolds numbers this model scale and Reynolds numbers may increase the friction drag
greater than 2.0 million but would have regions of transitional and values to the free transition levels of the 0.250 sub-scale model but it
laminar flow present at Reynolds numbers below 2.0 million. Results is unlikely that full-scale values can be replicated given the
for the 0.250 scale model show laminar and transitional flow below compressibility limitation of the 0.125 scale models.
1.0 million with transitional and small regions of turbulent flow
possible above 1.0 million. The 0.250 scale model is compressibility The ability to accurately represent the boundary layer state and to
limited to test Reynolds numbers below 4.0 million indicating that obtain representative full-scale friction drag in wind tunnel tests and
forced transition test methods should be used for all models at this computational simulations is critical if friction drag reduction is to be
scale to better simulate full-scale friction coefficient values. The realized. A number of friction drag reduction technologies have been
0.125 scale model results show compressibility effects limit testing to developed within the aircraft and watercraft communities that may be
Reynolds numbers below 2.0 million. The friction coefficient values applicable to commercial ground vehicles [343, 344, 345, 346, 347,
indicate that the boundary layer will be laminar over the majority of 348, 349, 350, 351, 352, 353, 354, 355, 356, 357, 358, 359, 360, 361,
the test model for all test Reynolds numbers. 362]. The following argument is provided in support of friction drag
reduction strategy.

If the drag of a future vehicle is evenly distributed between the tractor


and trailer, and the vehicle drag is 60 to 70 percent pressure drag and
the remainder friction drag [4], and trailer drag is evenly distributed
between pressure and friction drag, then a 50 percent reduction in
trailer friction drag will result in a 10 percent reduction in vehicle drag.

Figure 6.20. Variation in the average friction coefficient with Reynolds


number for various scale models of a combination vehicle.

The friction coefficient values shown in figure 6.20 were used to


calculate friction drag for each model scale as shown in figure 6.21.
For a full-scale vehicle with a drag coefficient of 0.40 the graph shows
that friction drag comprises 20 percent of the total drag at 1.0 million
Reynolds number and 16 percent of the total drag at 4.5 million Figure 6.21. Variation in the average friction drag coefficient with Reynolds
Reynolds number. A comparison of the drag for the friction drag number for various scale models of a combination vehicle.
coefficients for the full-scale, 0.250 scale and 0.125 scale test articles
show the 0.250 and 0.125 sub-scale models have friction drag A conceptual approach to reduce the friction drag on the trailer
coefficients well below full-scale values. The 0.250 scale results show follows:
drag levels that vary from 55 percent of full-scale at 1.0 million
Reynolds number to 70 percent of full-scale at 4.0 million Reynolds Trailer skin friction drag reduction may be accomplished by
number. As discussed in section 6.2 and depicted in figure 6.12 [177] geometric modifications to the trailer side and top surfaces and the
the use of surface roughness to promote early transition of the use of technologies that manage turbulent boundary layer structures.
boundary layer is a viable method to simulate higher Reynolds For example, the forward most area (laminar region) of each surface
number flow. The data of reference 175 indicate that surface roughness can be modified to optimize the extent of laminar flow by; reducing
can reduce the extent of both laminar and transitional boundary layer surface roughness and introducing minimal three-dimensional surface
and thereby would have increased the friction drag at all Reynolds shaping to impose a pressure gradient that will promote natural
numbers and may produce drag levels that approach the full-scale laminar flow. The mid section area of the trailer (transitional region)
friction drag at Reynolds numbers greater than 2.0 million. A review can then be modified to optimize the streamwise extent of transitional
of the 0.125 scale results show drag levels that vary from 33 percent boundary layer by; three dimensional shaping to manage spanwise
of full scale at 1.0 million Reynolds number to 45 percent of full-scale flows that promote desirable boundary thickness characteristics
at 2.0 million Reynolds number. Use of forced transition techniques at leading to the development of a preferred turbulent boundary layer.
The aft most area of each surface (turbulent region) can be modified
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 653

to control the structures within the turbulent boundary layer by; of the boundary layer conditions and pressure distribution on the
extending the three dimensional surface shaping strategy to manage vehicle. Proper use of this technology can improve the performance
boundary layer thickness and introducing streamwise structures that of an aerodynamic surface by 20 percent.
modify the turbulent boundary layer structures. The blending of
multiple boundary layer flow control strategies to achieve skin Passive porosity [375, 376, 377] reduces flow separation resulting
friction drag reduction would be applicable to low yaw conditions. from severe pressure gradients on the surface of a vehicle. The
technology is geometrically complex and consists of a porous surface
The friction drag reduction topic is typically divided into two efforts, placed over a minimal depth cavity in the region of a severe pressure
those focused on maintaining a laminar boundary layer and efforts gradient. The cavity allows communication between the high and low
reducing the turbulent boundary-layer friction drag. It is beyond the pressures eliminating the problematic pressure gradient. Application
scope of this paper to address the diversity and complexity of this of the technology is proposed for regions of abrupt geometric
topic. However, there are several examples worth noting. The first changes such as the attachment region of a boattail, the trailing edge
technology is riblets [345, 350, 355]. Riblets were conceived from of a boattail, or the junction between the hood and windshield. As
observations of nature [348, 349], specifically the skin of a shark, and with the VG technology the geometric design of the system is
they were originally developed for aircraft applications. They have Reynolds number dependent.
found their way onto watercraft and into pipes and ducts. It is
interesting to note that they have not achieved universal success in Base vents [372], boundary layer diverters and boundary layer
the aerospace community but a version of the technology is used in thickness control methods [370] are techniques that manage the
the Industrial area for pipe flows [350]. Other skin friction drag interaction of the boundary layer with a geometric discontinuity with
reduction technologies are laminar flow control with boundary layer the goal of reducing drag. The successful use of these techniques
removal [362] and compliant walls [352]. requires an understanding the boundary layer and its variability with
changes in Reynolds number and outside disturbances. Suggested
areas of application are upstream of regions of abrupt changes in
6.3.2. Flow Control
geometry, such as the trailing edge of a trailer.
Active and passive flow control technologies [34, 363, 364, 365, 366,
367] may be used to enhance the effectiveness, alter the effective Active systems include morphing surfaces, mass/temperature/energy
shape and reduce the wetted area of current and future fairings [160, (MTE) addition or subtraction systems, and oscillating/rotating/
176, 177, 181, 339, 340, 341, 342]. For most applications, flow flapping (ORF) structures, [378, 379, 380, 381, 382, 383, 384, 385,
control technologies would not replace a fairing. These technologies 386, 387, 388, 389, 390, 391]. In the area of MTE addition or
can be used to control flow separation at off-design points such as subtraction the work of Englar [378] is noteworthy for transferring
moderate and high yaw conditions or low speed operation where Coanda flow control technology from aircraft to ground vehicles. The
laminar boundary layers may dominate. Flow control systems can Coanda system employs a high velocity jet of air that is blown
also be used manage cooling flow in the presence of atmospheric and tangential to a curved surface located at the trailing edge of the trailer
vehicle interference. The difference between an active and passive side and top surfaces. The jet remains attached to the curved surface
system is a function of energy use. Contrary to an active system a and turns towards the trailer centerline. The flow turning induces the
passive technology does not require energy addition to generate the trailer boundary layer adjacent to the jet of air to also curve towards
desired effect. The difference between a flow control system and a the vehicle centerline. Application of this technology by Englar [378]
fairing is the ability of the flow control system to adapt to changing and Metka [392] indicate that operational vehicle fuel savings may
flow conditions while a fairing is a fixed structure that shields a high approach 5 percent.
drag region from the on-coming flow.
Recent active flow control efforts have focused on various types of
Passive systems include vortex generators, base vents, boundary layer synthetic jets, which are in the same category as the Coanda studies,
diverters, passive porosity and boundary layer thickness control see papers by Thomas [381], Kozlov [383], El-Alti [388] and Siefert
methods [368, 369, 370, 371, 372, 373, 374, 375, 376, 377]. Vortex [390]. Compared to the work of Englar the energy requirement for
generators (VG) [371, 372, 373, 374] are the most mature technology synthetic jets is significantly reduced however these systems are more
and have been accepted throughout the transportation and industrial complex. Both the Coanda and synthetic jet studies are focused on
communities. They can be found on aircraft, ground vehicles, base drag however unlike the Coanda studies the synthetic jet efforts
watercraft and in diffusers, ducts and heat exchangers. VGs generate do not intend to replace the boattail fairings but are attempting to
coherent vortex structures in boundary layers to minimize flow augment the performance of a boattail.
separation due to an adverse pressure gradient. VG have also been
designed as active devices and may be either structural or fluidic- Base bleed [384] is another MTE approach that may be viewed as an
based. Suggested use of this technology would be to suppress active version of the passive boundary layer thickness control
boundary layer separation on trailer roof fairings and trailer boattails technique [369]. Both methods add mass tot the base area in an
at low and moderate yaw conditions. While VGs may appear to be a attempt to reduce base pressures by controlling the level and
simple concept their geometric design and placement on a vehicle is distribution of vorticity in the trailing wake.
Reynolds number dependent and requires a thorough understanding
654 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

The final active flow control topic area is ORF structures as noted by Equally important in the discussion of platooning is to understand
a simple rotation of trailing base flap discussed by Johansson [380] or negative interference flows. Aerodynamic interference is present
the three dimensional morphing of a three dimensional structure between all vehicles operating on the road where the magnitude and
described by Barbarinao [391]. Both concepts are used in extensively reach of the interference is a function of the geometric features of the
in the aerospace community and have been used on ground vehicles vehicles involved.
for cooling flow management.
For heavy combination vehicles the following guidelines are
As with the passive systems the successful application of active suggested; 1) a leading vehicle will measurably influence a trailing
techniques requires an understanding of the boundary layer and its vehicle if it is within five body lengths, 2) a trailing vehicle will
variability with changes in Reynolds number and outside measurably influence a leading vehicle if it is within two body
disturbances. A foundational element of all active systems is the lengths, 3) laterally adjacent vehicles will influence each other if they
requirement that they are a “smart” system capable of changing their are within three body widths, and 4) if vehicles are laterally or
operational characteristics based on real time aerodynamic and flow longitudinally positioned within a half of a body width they may be
sensor feedback and not based on a global vehicle performance treated as a single vehicle. In general, longitudinal-only spacing will
parameter related to engine operation, vehicle speed or yaw angle. produce positive aerodynamic interference while lateral-only spacing
This level of complexity and sophistication will delay the will generate negative interference.
implementation of these systems until 2020 or later.
For the ideal case of a heavy truck platoon operating in still air and
on a road void of traffic the lead vehicle benefits from a positive
6.3.3. Atmospheric and Vehicle Interference Flows
pressure field exerted on its base area. Stagnation pressures acting on
Any vehicle on the highway must contend with a chaotic the front of the following vehicle create this pressure field. The
aerodynamic environment comprised of atmospheric and traffic trailing vehicle in a platoon benefits from operating in the turbulent
generated time dependent interfering flows that increase the wake left by all preceding vehicles. The turbulent wake trailing each
aerodynamic drag and fuel consumption for a combination vehicle. vehicle has a forward velocity that decays with time. The wake
The magnitude of this effect exceeds that experienced by any other forward velocity results in reduced pressure drag on the vehicle cab
vehicle on the road. The increase in aerodynamic drag is a result of a front and a local Reynolds number for the trailing vehicle that is
variation in flow separation along the length of the vehicle. lower than that experienced by the lead vehicle. The base drag of the
trailing vehicle will be different from a vehicle operating in still air
The time-dependent interfering flows will vary in magnitude and due to the change in local Reynolds number that will alter the
direction with a time scale that may be proportional to the vehicle boundary layer development along the vehicle length resulting in a
speed. This time dependent flow variability may disrupt the change in wake structure at the base. Following vehicles are located
stabilization of the vehicle boundary layer and flow separation between the lead and trailing vehicles. These vehicles benefit from
patterns introducing additional time dependencies into the flow o the both the positive pressure field acting on their base and operating in
surface of the vehicle. These effects are especially problematic for a the turbulent wake left by a preceding vehicle. Each subsequent
commercial vehicle that operates in a narrow window of economic following vehicle will experience a lower drag on the vehicle front
viability. Additionally these aerodynamic interference flows may face (reduced pressure drag on the cab front) as a result of operating
mask projected fuel savings benefits that are based on controlled in an increasing turbulent wake but will realize a reduced benefit
aerodynamic and/or fuel consumption tests. from the pressure field exerted on its base as a result of the reduced
pressure drag on the cab front of its following vehicle. As mentioned
Accounting for the influence of atmospheric and vehicle interference previously these changes in interference effects result from the
flows in the design process is required to maximize vehicle reduction in local Reynolds number between successive vehicles in a
performance. Success in this area may require tractor and trailer platoon. This local Reynolds number will impact the aerodynamic
multi-point design shaping coupled with boundary layer management performance of each vehicle compared to still air operations.
and active flow control. The large number of variables in this problem Including real world factors, such as yaw and traffic generated
set is beyond the scope of this section. To provide a focus for this unsteady flow features would alter platoon performance and
topic the following content will look at a subset of this complex introduce additional Reynolds number effects.
problem, platooning of heavy combination vehicles. For this
discussion a platoon is comprised of a lead vehicle and one or more Recent experimental and computational studies of platoons
following vehicles where the last vehicle in a platoon will be labeled comprised of heavy combination vehicles or generic representations
the trailing vehicle, of heavy vehicles [5, 173, 392, 393, 394, 395, 396, 397, 398, 399,
400, 401, 402, 403, 404, 405] suggest fuel savings between 5 percent
Platooning has been a topic of interest since the 1960s and over the and 20 percent, see figure 6.21. The large spread in results is a
past few years has seen increased activity tied to the promise of product of the variability in vehicles/geometries, platoon
significant drag reductions and associate fuel savings, compared to a characteristics and test/analysis tools used.
single vehicle. Platoon drag reduction is the product of positive
aerodynamic interference between longitudinally aligned vehicles.
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 655

results from a combination of additional tractor aerodynamic


shaping and the addition of aerodynamic fairings to the tractor and
trailer. The ability to achieve the additional drag reduction beyond
2015 will be dependent on the management of both pressure and
friction drag forces. One byproduct of the pressure drag reduction
is the growing importance of friction drag. Published data suggest
that the industry average drag coefficient of next generation
in-service combination vehicles will be less than 0.50 with a
potential to achieve 0.40 by 2030.

Reynolds number investigations related to commercial vehicle drag


reduction began with the NASA studies in the 1970s. By 1979 SAE
Figure 6.21. Photograph of road test platoon of heavy combination vehicles [392].
published Recommended Practice J1252 entitled “Wind Tunnel Test
A finding consistent for two 2014 studies [399, 400] is the loss in Procedure for Trucks and Buses”. This document provided the first
cooling performance for a following vehicle in a platoon. set of wind tunnel testing criteria for commercial vehicles and
Computational simulations by Smith [399] show that following established the original set of Reynolds number guidelines and
vehicles will see a 40 percent drop in mass flow through the heat proposed a minimum test Reynolds number of 0.7 million. In 1981
exchanger and Lammert [400] shows a loss in fuel savings from Gilhaus proposed 1.6 million [35], Cooper proposed 2.0 million in
8.4 percent to 2.8 percent due to cooling fan operation to 1982 [23] and 1984 [330], Drollinger suggested 2.0 million in 1987
compensate for a loss in ram air. A possible contributing factor to [1], Olson offered 1.7 million in 1992 [46] and in 1998 Hucho
the drop in cooling flow may be the drop in local Reynolds number suggested 2.5 million [40]. Contributing to the body of evidence
at the front face of the tractor. A potential solution to the cooling supporting an increase in the test Reynolds number were several
flow matter would be the use of active flow control systems to contributions in the 2000 to 2010 period. In 2004 Storms [54] data
manage conflicting ram air requirements between still air and suggested 3.0 million and Hjelm's paper [197] showed 4.0 million.
traffic/platoon operation. It should be noted that these studies were In 2012 the 0.7 million Reynolds number was removed from J1252
performed with current model vehicles with engine displacements and replaced with a detailed test procedure to determine the
larger than that required for low drag next generation vehicles. minimum allowable Reynolds number for the specific test being
These future vehicles would have lower cooling flow requirements conducted. The intent of the test Reynolds number assessment
that may not be adversely affected by a leading vehicle. procedure is to allow adjustments to the minimum test Reynolds
number based on differences in the test vehicle or the specific
Further investigations of aerodynamics and/or fuel consumption operational Reynolds number range of the reference full-scale
impacted by atmospheric, traffic and platoon interference requires vehicle. Contributing to the body of work is the 2012 paper by Wood
significantly greater resources than those for a single vehicle. Flow [2] that proposed 3.0 million, Leuschen [65] recommends 4.0
similarity requirement eliminates wind tunnel as a viable tool and million McArhthur [66] shows a value greater than 2.4 million and
imposes a significant burden on both computational simulations and the DOE published full-scale data [181] that shows greater than 2.0
operational testing. For example, computational simulations may million. Competing with the published Reynolds number data are
require new boundary layer models to capture changes in local EPA and CARB regulatory criteria that sets the minimum test
Reynolds numbers and boundary layer states. Operational testing will Reynolds number at 1.0 million.
require extensive use of instrumentation to capture aerodynamic data
to support assessment of variations in local Reynolds number and This Chapter has shown how Reynolds number impacts both data
boundary layer state. trends and magnitudes and that test or analysis results obtained at less
than full-scale Reynolds numbers will differ from drag levels at
full-scale Reynolds number by more than ten percent. The impact of
6.4. Summary Findings Reynolds number on vehicle aerodynamics is very complex and will
Over the past four decades the aerodynamic drag coefficient of the vary with vehicle shape, atmospheric conditions, aerodynamic
majority of operational heavy combination vehicle has been treatments, and both vehicle speed and yaw angle. To address these
reduced by more than fifty percent through the use of established critical factors the SAE has revised J1252 and J1321 and created
drag reduction technology and computational and experimental J2966. The SAE has activity underway to revise J1526 and develop
methods in the design process. By 1990 aerodynamic design both coastdown and constant speed standards specifically designed to
concepts, driven by the national 55mph speed limit and the 1970 address the unique test needs of commercial vehicles. However, the
energy crisis, had gained a foothold in the industry, truck cabs were actions of the regulatory community highlight the need for additional
undergoing gradual evolution to more aerodynamic shaping, edge refinements to the SAE commercial vehicle standards to clarify the
rounding and tractor roof fairings became the norm and importance and impact of Reynolds number effects on aerodynamic
computational methods were being used in the design process. and fuel consumption characteristics.
These modest aerodynamic improvements pushed drag coefficients
below 0.70 by 2000. The decrease in drag from 2000 to 2015
656 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

To achieve drag reduction beyond 2015 will require a system level operational testing will require extensive use of instrumentation to
perspective that requires aerodynamic improvements are balanced capture aerodynamic data to support assessment of variations in local
against all vehicle freight efficiency concepts. Near term Reynolds number and boundary layer state.
opportunities to maximize the benefit of current low drag vehicle
systems are; Critical to the success of each topic area is the continued
development of aerodynamic test and analysis tools capable of
1. Reclassify container carriers as trailers and subject them to all capturing the Reynolds number sensitivities present with future
the regulations of van trailers, designs. Of particular importance is the maturation of computational
2. Revise legislation to expand use of long combination vehicles simulation tools with improved friction modeling capability
(LCV), specifically developed for commercial ground vehicles.6.5.
3. Revise legislation to permit longer trailers,
4. Mandated speed limiters set to 60 mph or some low value. All 6.5. References
of these effectively reduce the fuel consumption of comparable 323. Mason, W. and Beebe, P., “The Drag Related Flow Field Characteristics
current vehicles through net drag force reduction, less fuel of Trucks and Buses,” Aerodynamic Drag Mechanisms of Bluff Bodies
and Road Vehicles. Symposium Held at the General Motors Research
required to move comparable freight loads. Laboratories, Plenum Press, 1978.
324. Hucho, W. and Sovran, G., “Aerodynamics of Road Vehicles,” Annual
Aerodynamic improvements for future vehicles may require friction Review of Fluid Mechanics, 25(1): 485-537, 1993.
325. Choi, H., Lee, J. and Park, H., “Aerodynamics of heavy Vehicles,”
drag reduction concepts, flow control systems, and management of Annual Review of Fluid Mechanics 46:441-468, doi:10.1146/annurev-
atmospheric and vehicle interference flows. Friction drag reduction fluid-011212-140616, 2014.
concepts can be used to reduce the turbulent friction coefficient and 326. Wood, R. M., “Aerodynamic Drag and Drag Reduction: Energy and
Energy Savings,” AIAA 41st Aerospace Sciences Meeting and Exhibit.
promote increased laminar flow by suppressing boundary layer AIAA - 2003-0209, Reno, NV, Invited Lecture, AIAA 2003-0209, Jan.
transition. Friction drag may also be reduced in a synergistic design 2003 Invited Lecture
process where flow control systems are used to improve fairing 327. “MOVES2010 Highway Vehicles Population and Activity Data,”
Assessment and Standards Division, Office of Transportation and Air
performance allowing for the minimization of fairing surface area Quality, U.S. Environmental Protection Agency, EPA-420-R-10-026,
that further reduce friction drag. Pursuing these advanced flow 2010.
management concepts requires an understanding of the boundary 328. Salvadori, S., Morbiato, T., Mattana, A. and Fusto, E., “On the
Characterization of Wind Profiles Generated by Road Traffic,” The
layer flows present on a commercial vehicle and any variations in Seventh International Colloquium on Bluff Body Aerodynamics and
friction drag and limitations associated with sub-scale wind tunnel Applications (BBAA7) Shanghai, China, International Association for
testing. Active and passive flow control technologies may be used to Wind Engineering, Sept. 2012.
329. Mayfield, D., “Turning Flights of Fancy into Reality,” The Virginia Pilot
enhance the effectiveness, alter the effective shape and reduce the and Ledger Star, Hampton Roads Business Weekly. Pg. 16-17. February
wetted area of current and future fairings. However, flow control 2, 1987.
technologies cannot and should not replace a fairing. These 330. Cooper, K., “Wind-Tunnel Simulation of Surface Vehicles,” Journal of
Wind Engineering and Industrial Aerodynamics, 17 (1984) 167-198
technologies can be used to control flow separation at off-design
331. “Recovery Act - Systems Level Technology Development, Integration,
points such as moderate and high yaw conditions or low speed and Demonstration for Efficient Class 8 Trucks (SuperTruck) and
operation where laminar boundary layers may dominate. Flow control Advanced Technology Powertrains for Light Duty Vehicles (ATP-
LD),” Funding opportunity announcement DE-FOA-0000079, National
systems can also be used manage cooling flow in the presence of Energy Technology Laboratory, U.S. Department of Energy, http://www.
atmospheric and vehicle interference. Interference flows were stateenergyoffice.wi.gov/docview.asp?docid=17004&locid=160, July
discussed in the context of platooning. Time-dependent interfering 2009. documents/zeDE-FOA-0000079.pdf.
332. Committee to Review the 21st Century Truck Partnership, Phase 2,
flows will vary in magnitude and direction with a time scale that may “Review of the 21st Century Truck Partnership, Second Report,”
be proportional to the vehicle speed. This time dependent flow National Research Council, The National Academies Press, Washington,
variability may disrupt the stabilization of the vehicle boundary layer D.C., ISBN 978-0-309-22247-1, 2012.
333. “Greenhouse Gas Emissions Standards and Fuel Efficiency Standards
and flow separation patterns introducing additional time dependencies for Medium- and Heavy-Duty Engines and Vehicles,” Final Rule,
into the flow on the surface of the vehicle. These effects are Environmental Protection Agency and National Highway Traffic and
especially problematic for a commercial vehicle that operates in a Safety Administration, Department of Transportation, Federal Register
76(179): 57106, September 15, 2011.
narrow window of economic viability. Additionally these 334. DOE Updated 21st Century Truck Partnership Roadmap and Technical
aerodynamic interference flows may mask projected fuel savings White Papers. Doc. No. 21CTP-003. September 1. Washington, D.C.:
benefits that are based on controlled aerodynamic and/or fuel Office of FreedomCAR and Vehicle Technologies.
335. Review of the 21st Century Truck Partnership. Washington, D.C.: The
consumption tests. Further investigations of aerodynamics and/or fuel National Academies Press, National Research Council, 2008.
consumption impacted by atmospheric, traffic and platoon 336. “Interim Test Method for Verifying Fuel-Saving Components for
interference requires significantly greater resources than those for a SmartWay: Modifications to SAE J1321,” U.S. Environmental
Protection Agency, National Service Center for Environmental
single vehicle. Flow similarity requirement eliminates wind tunnel as Publications, EPA-420-F-09-046, 2011.
a viable tool and imposes a significant burden on both computational 337. Cooper, K., “A Wind Tunnel Investigation into the Fuel Savings
simulations and operational testing. For example, computational Available from the Aerodynamic Drag Reduction of Trucks,” National
Research Council Canada, Department of Mechanical Engineering /
simulations may require new boundary layer models to capture National Aeronautical Establishment, Quarterly Bulletin No. 3, 1976.
changes in local Reynolds number and boundary layer states and
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 657

338. Håkansson, C. and Lenngren, M., “CFD Analysis of Aerodynamic 362. Hwang, D., “A Proof of Concept Experiment for Reducing Skin Friction
Trailer Devices for Drag Reduction of Heavy Duty Trucks,” Master's by Using a Micro-Blowing Technique,” 35th AIAA Aerospace Sciences
Thesis, Chalmers University of Technology, Department of applied Meeting and Exhibit, Reno, NV, AIAA 97-0546, Jan. 1997.
Mechanics, Division of Vehicle Engineering and Autonomous Systems, 363. Wood, R., “Impact of Advanced Aerodynamic Technology on
Göteborg, Sweden, 2010. Transportation Energy Consumption,” SAE Technical Paper 2004-01-
339. Wood, R., “EPA Smartway Verification of Trailer Undercarriage 1306, 2004, doi:10.4271/2004-01-1306.
Advanced Aerodynamic Drag Reduction Technology,” SAE Int. J. 364. Jahanmiri, M., “Active Flow Control: A Review,” Chalmers University
Commer. Veh. 5(2):607-615, 2012, doi:10.4271/2012-01-2043. of Technology, Göteborg, Sweden, Department of Applied Mechanics,
340. Wood, R., “Operationally-Practical & Aerodynamically-Robust Heavy Division of Fluid Dynamics, Research report 2010:12, ISSN 1652-8549,
Truck Trailer Drag Reduction Technology,” SAE Int. J. Commer. Veh. 2010.
1(1):237-247, 2009, doi:10.4271/2008-01-2603. 365. Gad-el-Hak, M., Pollard, A., and Bonnet, J., Eds. “Flow Control:
341. Cary, W., Landman, D., and Wood, R., “Experimental Investigation of Fundamentals and Practices,” Lecture Notes in Physics Monographs
Wake Boards for Drag Reduction on an Ahmed Body,” SAE Technical M53, Springer-Verlag, Berlin, Germany, ISBN: 3-540-63936-5, 1998.
Paper 2006-01-3530, 2006, doi:10.4271/2006-01-3530. 366. Kral, L., “Active Flow Control Technology,” ASME Fluids Engineering
342. Wood, R. and Bauer, S., “Simple and Low-Cost Aerodynamic Drag Division Technical Brief, 2000
Reduction Devices for Tractor-Trailer Trucks,” SAE Technical Paper 367. Lachman G, editor. “Boundary Layer and Flow Control,” vols. 1 and 2.
2003-01-3377, 2003, doi:10.4271/2003-01-3377. Oxford: Pergamon Press; 1961.
343. Bushnell, D. and Moore, K., “Drag Reduction in Nature,” Annual 368. Gustavsson, T., “Alternative Approaches to Rear End Drag Reduction,”
Review of Fluid Mechanics, 23: 65-79, 1991. Thesis, KTH Royal Institute of Technology, Department of Aeronautical
344. Bushnell, D., “Longitudinal Vortex Control-Techniques and and Vehicle Engineering, TRITA-AVE 2006:12, ISSN: 1651-7660, 2006.
Applications,” The 32nd Lanchester Memorial Lecture, Royal 369. Howard, F. and Goodman, W., “Axisymmetric Bluff-Body Drag
Aeronautical Society, London, Aeronautical Journal 96(958):293-312, Reduction Through Geometrical Modification,” AIAA Journal of
October 1992. Aircraft. 22(6): 516-522, 1985.
345. Bushnell, D. and Hefner, J., Eds., “Viscous Drag Reduction in Boundary 370. Tanner, M., “Boundary-layer Thickness and Base Pressure,” AIAA
Layers,” Progress in Astronautics and Aeronautics 123, American Journal 23(12): 1987-1989, 1985.
Institute of Aeronautics and Astronautics, Reston, VA, ISBN: 978-0-
371. Grosche, F. and Meier, G., “Research at DLR Gottingen on Bluff Bodies
930403-66-9, 1990.
Aerodynamics, Drag Reduction by Wake Ventilation and Flow Control,”
346. Farrell, B. and Ioannou, P., “Turbulence Suppression by Active Control,” Journal of Wind Eng. and Ind. Aerodynamics 89(14-15): 1201-1218,
Physics of Fluids 8(5):1257-1268, 1996. 2001.
347. Gad-el-Hak, M., “Interactive Control of Turbulent Boundary Layers: A 372. Viswanth, P., “Flow Management Techniques for Base and Afterbody
Futuristic Overview,” AIAA Journal 32(9):1753-1765, 1994. Drag Reduction,” Progress in Aerospace Sciences 32(2-3):79-129. 1996.
348. Dean, B. and Bhushan, B., “Shark-Skin Surfaces for Fluid-Drag 373. Lin, J., Selby, G., and Howard, F., “Exploratory Study of Vortex-
Reduction in a Turbulent Flow: A Review,” Philosophical Transactions generating Devices for Turbulent Flow Separation Control,” 29th AIAA
of the Royal Society A 368(1929): 4775-4806, 2010. Aerospace Sciences Meeting, Reno, NV, Paper 42, 1991.
349. Lynch, F. and Klinge, M., “Some Practical Aspects of Viscous 374. Lin, J., “Control of Turbulent Boundary Layer Separation Using Micro-
Drag Reduction Concepts,” SAE Technical Paper 912129, 1991, Vortex Generators,” 30th AIAA Fluid Dynamics Conference, Norfolk,
doi:10.4271/912129. VA, AIAA 99-3404, 1999.
350. Christodoulou, C., Liu, K., and Joseph, D., “Combined Effects of Riblets 375. Wood, R., Banks, D., and Bauer, S., “Passive Porosity with Free and
and Polymers on Drag Reduction in Pipes,” Physics of Fluids A: Fluid Fixed Separation on a Tangent-Ogive Forebody,” AIAA Journal of
Dynamics 3(5): 995-996, 1991. Aircraft 31(5): 1219-1221, 1994.
351. Kramer, B., Smith, B., Heid, J., Noffz, G., Richwine, D., and Ng, T., 376. Hunter, C., Viken, S., Wood, R., and Bauer, S., “Advanced Aerodynamic
“Drag Reduction Experiments Using Boundary Layer Heating,” 37th Design of Passive Porosity Control Effectors,” AIAA 2001-0249, 39th
Aerospace Sciences Meeting and Exhibit, Reno, NV, AIAA-99-0134. AIAA Aerospace Sciences Meeting, Reno NV, Jan. 2001.
Jan. 1999.
377. Wood, R., “A Discussion of Aerodynamic Control Effectors (ACEs)
352. Lekoudis, S. and Sengupta, T., “Two-Dimensional Turbulent Boundary for Unmanned Air vehicles (UAVs),” AIAA 2002-3494, 1st AIAA UAV
Layers Over Rigid and Moving Swept Wavy Surfaces,” Physics Fluids Conference Portsmouth, VA, May 2002.
29(4): 964-970, 1986.
378. Englar, R., “Improved Pneumatic Aerodynamics for Drag Reduction,
353. Sahlin, A., Alfredsson, P. H., Johansson, A., “Direct Drag Measurements Fuel Economy, Safety And Stability Increase for Heavy Vehicles,” SAE
for a Flat Plate with Passive Boundary Layer Manipulators,” Physics Technical Paper 2005-01-3627, doi:10.4271/2005-01-3627.
Fluids 29(3): 696-700, 1986.
379. Van Leeuwen, P., “Computational Analysis of Base Drag Reduction
354. Lee, C and Kim, J., “Control of the Viscous Sublayer for Drag Using Active Flow Control,” Master's Thesis, Delft University of
Reduction,” Physics of Fluids, 14(7): 2523-2529, 2002. Technology, The Netherlands, Faculty of Aerospace Engineering, 2009.
355. Baron, A. and Quadrio, M., “Turbulent Boundary Layer Over Riblets: 380. Johansson, M., “The Effects of Oscillatory Boat-Tail Flaps on the
Conditional Analysis of Ejection-Like Events,” International Journal of Near Wake of a Prototypical Heavy Vehicle,” Master's Thesis, Luleå
Heat and Fluid Flow 18(2):188-196, 1997. University of Technology, Sweden, Department of Applied Physics and
356. Kerho, M., “Active Reduction of Skin Friction Drag Using Low-Speed Mechanical Engineering / Physics, 2010
Streak Control,” 40th AIAA Aerospace Sciences Meeting and Exhibit, 381. Thomas, F., Kozlov, A., and Corke, T., “Plasma Actuators for Bluff Body
Reno, NV, AIAA 2002-0271, Jan. 2002. Flow Control,” AIAA 2006-2845, 3rd AIAA Flow Control Conference,
357. Choi, K., Yang, X., Clayton, B., Glover, E., Atlar, M., Semenov, B., San Francisco, CA, June 2006.
and Kulik, V., “Turbulent Drag Reduction Using Compliant Surfaces,” 382. Koslov, A., “Plasma Actuators for Bluff Body Flow Control,” Ph.D.
Proceedings of the Royal Society London A 453(1965): 2229-2240, Thesis, University of Notre Dame, IN, URN: etd-07232010-105634,
1997. 2010.
358. Ashill, P., Fulker, J. and Hackett, K., “Studies of Flows Induced by Sub 383. Kozlov, A., Thomas, F., “Active Noise Control of Bluff-Body Flows
Boundary Layer Vortex Generators (SBVGs),” 40th AIAA Aerospace Using Dielectric Barrier Discharge Plasma Actuators,” AIAA paper
Sciences Meeting and Exhibit, Reno, NV, AIAA-2002-0968, Jan. 2002. 2009-3245, Proceedings of the 15th AIAA/CEAS Aeroacoustics
359. Walsh, M., “Riblets as a Viscous Drag Reduction Technique,” AIAA Conference, 2009.
Journal 21(4): 485-486, 1983. 384. Bearman, P., “The Effect of Base Bleed on the Flow Behind a Two-
360. Neiuwstadt, F., Wolthers, W., Leijdens, H., Krishna Prasad, K. and Dimensional Model with a Blunt Trailing Edge,” Royal Aeronautical
Schwarz-van Manen, A., “The Reduction of Skin Friction by Riblets Society, Aeronautical Quarterly 18:207-224, 1967.
Under the Influence of an Adverse Pressure Gradient,” Experiments in 385. Manosalvas, D., Economon, T., Palacios, F. and Jameson, A., “Finding
Fluids 15(1):17-26, 1993. Computationally Inexpensive Methods to Model the Flow Past Heavy
361. Hefner, J. and Bushnell, D., “Viscous Drag Reduction Via Surface Mass Vehicles and the Design of Active Flow Control Systems fro Drag
Injection,” chapter in source [345] Reduction,” AIAA 2014-2404, 32nd AIAA Applied Aerodynamics
Conference, Atlanta, GA, June, 2014.
658 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

386. El-Alti, M., Chernoray, V., Kjellgren, P., Hjelm, L., and Davidson, L., 395. Browand, F., McArthur, J., and Radovich, C., “Fuel Saving Achieved
“Computations and Full-scale tests of Active Flow Control Applied in the Field Test of Two Tandem Trucks,” California PATH Research
on a VOLVO Truck-Trailer,” Engineering Conferences International, Report UCB-ITS-PRR-2004-20, University of California, 2004.
Aerodynamics of Heavy Vehicles III: Trucks, Buses and Trains, 396. Pagliarella, R., Watkins, S., and Tempia, A., “Aerodynamic Performance
Potsdam, Germany, Sept. 2010. Available in:Dillmann, A. and Orellano, of Vehicles in Platoons: The Influence of Backlight Angles,” SAE
A., Eds., “The Aerodynamics of Heavy Vehicles III: Trucks, Buses and Technical Paper 2007-01-1547, 2007, doi:10.4271/2007-01-1547.
Trains,” Springer International Publishing, ISBN: 978-3-319-20121-4,
397. Al Alam, al A., Gattami, A., and Johansson, K., “An Experimental Study
2015.
on the Fuel Reduction Potential of Heavy Duty Vehicle Platooning,”
387. El-Alti, M., Chernoray, V., Kjellgren, P., Jahanmiri, M., and Davidson, 2010 13th International IEEE Annual Conference on Intelligent
L., “Experimental and Computational Studies of Active Flow Control Transportation Systems, September 19-22, 2010.
on a Model Truck-Trailer,” EFM 11: Experimental Fluid Mechanics
398. Tsugawa, S., Kato, S., and Aoki, K., “An Automated Truck Platoon
2011, Edited by Tomáš Vít, Petra Dančová and Petr Novotn Technical
for Energy Saving,” 13th International IEEE Conference on Intelligent
University of Liberec, Proceedings of the International Conference, Vol.
Transportation Systems, Funchal, Portugal, Sept. 2010.
2. pp. 600-615, 2011
399. Smith, J., Mihelic, R., Gifford, B., and Ellis, M., “Aerodynamic Impact
388. Zhang, B., Zhou, Y. and To, S., “Active Drag Reduction of a Simplified of Tractor-Trailer in Drafting Configuration,” SAE Int. J. Commer. Veh.
Car Model Using a Combination of Steady Actuations,” 19th 7(2):619-625, 2014, doi:10.4271/2014-01-2436.
Australasian Fluid Mechanics Conference Melbourne, Australia 8-11
December 2014 400. Lammert, M., Duran, A., Diez, J., Burton, K. et al., “Effect of Platooning
on Fuel Consumption of Class 8 Vehicles Over a Range of Speeds,
389. Sieifert, A., Stalnov, O., Sperber, D., Arwatz, G., Palei. V., David, S., Following Distances, and Mass,” SAE Int. J. Commer. Veh. 7(2): 626-
Dayan, I., and Fono, I., “Large Trucks Drag Reduction Using Active 639, 2014, doi:10.4271/2014-01-2438.
Flow Control,” AIAA 2008-743, 46th AIAA Aerospace Sciences Meeting
and Exhibit, Reno, NV, Jan. 2008. Also available in:The Aerodynamics 401. Vegendia, P., Sofu, T., Saha, R., Kumar, M. et al., “Investigation of
of Heavy Vehicles II: Trucks, Buses, and Trains, Lecture Notes in Aerodynamic Influence on Truck Platooning,” SAE Technical Paper
Applied and Computational Mechanics 41:115-133, 2009. 2015-01-2895, 2015, doi:10.4271/2015-01-2895.
390. Metka, M., “An Examination of Active Drag Reduction Methods for 402. Mihelic, R., Smith, J., and Ellis, M., “Aerodynamic Comparison of
Ground Vehicles,” Undergraduate Honors Thesis, Ohio State University, Tractor-Trailer Platooning and A-Train Configuration,” SAE Int. J.
Department of Mechanical and Aerospace Engineering, 2013. Commer. Veh. 8(2):740-746, 2015, doi:10.4271/2015-01-2897.
391. Barbarinao, S., Bilgen, O., Ajaj, R, Friswell, M. and Inman, D.:“A 403. Ellis, M., Gargoloff, J., and Sengupta, R., “Aerodynamic Drag and
Review of Morphing Aircraft,” Journal of Intelligent Material Systems Engine Cooling Effects on Class 8 Trucks in Platooning Configurations,”
And Structures 22(9): 823-877, doi:10.1177/1045389X11414084, 2011. SAE Int. J. Commer. Veh. 8(2):732-739, 2015, doi:10.4271/2015-01-
2896.
392. Doppenberg, S., “Drag Influence of Tails in a Platoon of Bluff Bodies,”
Master's Thesis 024#15#MT#FPP, Delft University of Technology, The 404. Koenig, K., “Interference Effects on Drag of Bluff Bodies in
Netherlands, Faculty of Aerospace Engineering, Department of Flight Tandem,” Ph.D. Thesis, California Institute of Technology, Division of
Performance and Propulsion, 2015. Engineering and Applied Science, 1978.
393. Bonnet, C. and Fritz, H., “Fuel Consumption Reduction in a Platoon: 405. Rajamani, G., “CFD Analysis of Air Flow Interactions in Vehicle
Experimental Results with two Electronically Coupled Trucks at Close Platoons,” Master's Thesis, RMIT University, Melbourne, Australia,
Spacing,” SAE Technical Paper 2000-01-3056, 2000, doi:10.4271/2000- Department of Aerospace, Mechanical and Manufacturing Engineering,
01-3056. 2006.
394. Hammache, M., Michaelian, M. and Browand, F., “Aerodynamic Forces
on Truck Models Including Two Trucks in Tandem,” California PATH
Research Report, UCB-ITS-PRR-2001-27, 2001.
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 659

CHAPTER 7. CONCLUDING REMARKS A discussion of computational simulation for commercial vehicle


design and analysis shows that the emergence of CFD in commercial
The commercial trucking industry is the keystone of our economy but
vehicle aerodynamic studies began in the 1990s. This growth has
continues to be burdened by the cost of fuel, the most volatile
been fueled by the reduced cost, shorter solution timeline, increased
operational expense. Complicating the equation are the growing
accessibility and portability of results for CFD, compared to wind
environmental-based regulatory requirements that are driving up
tunnel testing. The influence of CFD on the commercial vehicle
vehicle cost and maintenance demands. Both the industry and
industry extends well beyond external aerodynamics and includes
regulators have recognized aerodynamic drag reduction as a viable
internal flows, combustion, hydraulics, and thermal management. In
solution to these challenges. The majority of effort to date has
each of these areas Reynolds number plays a key role in defining the
focused on aerodynamic fairing concepts where test data suggest
fluid dynamics and in guiding method application. However setting
reductions in aerodynamic drag up to 40 percent may be possible
Reynolds number in the computational simulation does not, by itself,
with corresponding improvements in fuel economy approaching 20
result in flow similarity. Small geometric details, free stream
percent. However these projected improvements have not been
turbulence levels, surface roughness, the chosen turbulence modeling
consistently realized under operational conditions due to the narrow
and the size of the grid cells as well as the grid refinement in the
design window of the aerodynamic system and the significant
boundary layer are sensitive parameters with high impact on the
difference between the operational and test environment.
boundary layer state. An error leading to a delayed boundary layer
transition, for instance, can have influence on flow separation,
To minimize the uncertainty in vehicle performance the aerodynamic
changing the whole flow field around a truck.
process must address off-design operating performance and take into
account environmental factors, such as wind disturbances from traffic
When accurate prediction of separated wake flows and the impact on
and weather. Compared to the ideal test case these factors can
aerodynamic drag are needed, CFD methods which capture the
dominate the aerodynamic drag of commercial vehicles. In addition
turbulent, transient flow structures in separated regions using the
the freight industry requires aerodynamic solutions that are
highest possible fidelity should be used. Methods such as VLES or
operationally practical, aerodynamically robust, and satisfy all safety
DES are recommended over RANS so that turbulent flow structures
requirements. It is within this context that Reynolds number and
are simulated rather than modeled. The use of RANS and URANS
boundary layer sensitivity factors will influence aerodynamic design
should be limited to preliminary design and analysis of vehicle/
and vehicle performance.
models components with attached flow and known boundary layer
characteristics. HYBRID and LES methods should be used, with
The concept of dimensional analysis and flow (aerodynamic)
experimentally determined boundary layer characteristics, for
similarity was discussed to show the aerodynamics of bluff body
detailed design vehicles/models, aerodynamic components and
commercial ground vehicles is only dependent on the Reynolds
evaluation of vehicle aerodynamic performance. Obtaining a relevant
number similarity parameter. Bluff body aerodynamic trends were
solution requires a basic understanding of fluid dynamics and
reviewed and discussed in increasing levels of complexity from
computational sciences. To obtain the best available solution requires
simple two- and three-dimensional shapes, to major vehicle
an ability to make expert judgments in fluid dynamics and
components and finally related to sub-scale and full-scale vehicles.
computational sciences.
Wind tunnel data served as the source for all discussions. The
terminology of Roshko was introduced and adopted for discussing
A review of capabilities and limitations of experimental test methods
that the variation in drag with Reynolds number in which the
employed to evaluate the aerodynamics of commercial vehicles
subcritical, critical, transitional and transcritical flow regimes were
included wind tunnel test, coastdown, constant speed, and fuel
defined. Also discussed was the impact of the three boundary layer
consumption testing. While wind tunnel testing is the only method that
states (subcritical, transitional and transcritical) on the aerodynamics
obtains aerodynamic force and moment data all methods are
of this vehicle class.
susceptible to Reynolds number effects and must satisfy geometric and
flow similarity requirements. The importance of performing
Data for simple two- and three-dimensional shapes show that the
aerodynamic testing at Reynolds number values in which the boundary
transcritical Reynolds number can vary over a broad Reynolds
layer characteristics and aerodynamic forces are similar to those of the
number range for small changes to a vehicle component. In contrast,
full-scale vehicle at operational conditions were discussed along with
vehicle component data as represented by edge radius, boattail and
the need to obtain measurement of the boundary layer characteristics of
wheels all showed a transcritical Reynolds number greater than 2
the full-scale vehicle at operational condition. To guide commercial
million. A summary of the vehicle data identifies the minimum
vehicle testing the SAE has established recommended practice J1252
transcritical Reynolds number as 3 million. The vehicle data also
for wind tunnel testing and for fuel consumption testing the two
highlighted the complexity and Reynolds number dependence of the
primary methods are J1321 and J1526. While the SAE recognizes the
flow over a commercial vehicle and clarified the importance of
need for commercial vehicle standards to support coastdown and
modeling the full-scale aerodynamics as well as on-body and
constant speed testing they remain in development.
off-body flow conditions when conducting experimental and
computational studies at sub-scale or off-design conditions.
The importance of Reynolds number similarity to wind tunnel testing
was discussed in the context of producing results representative of the
operational vehicle. Recognizing that Reynolds similarity is complex
and dependent on many factors a series of criteria were provided for
660 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

the comparison of wind tunnel to operational vehicles, operational full-scale vehicle. Contributing to the body of work is the 2012 paper
vehicle components and for comparing results between wind tunnel by Wood that proposed 3.0 million, Leuschen recommends 4.0
tests. To provide context an overview of Reynolds number sensitivities million McArthur shows a value greater than 2.4 million and the
with wind tunnel facility type, floor/road simulation options, model- DOE published full-scale data that shows greater than 2.0 million.
support considerations and with yaw testing has been given. Competing with the published Reynolds number data are EPA and
CARB regulatory criteria that sets the minimum test Reynolds
Road load testing options and guidelines were presented that indicate number at 1.0 million. These actions of the regulatory community
coastdown testing of commercial vehicles should require vehicle highlight the need for additional refinements to the SAE commercial
speeds to stay above 30mph and restrict wind speed to less than 5 vehicle standards to clarify the importance and impact of Reynolds
mph. Underlying concerns were outlined relative to Reynolds number number effects on aerodynamic and fuel consumption characteristics.
state and the importance to tailor the coastdown and constant speed
test conditions to fall within the criteria for the specific Reynolds To achieve drag reduction beyond 2015 will require an all-inclusive
number state. This requires that the Reynolds number state of the test vehicle system level design perspective rather than summing
vehicle at operational conditions of interest shall be determined prior individual technology contributions. Aerodynamic drag reduction will
to testing. The following guidance is offered based on the vehicle require the use of friction drag reduction concepts, flow control
operational conditions of interest; 1) for subcritical Reynolds number systems, and management of atmospheric and vehicle interference
range operations the test shall not be constrained by Reynolds flows. Friction drag reduction concepts can be used to reduce the
number 2) for transitional Reynolds number range operations a test turbulent friction coefficient and promote increased laminar flow by
shall not be performed, and 3) for transcritical Reynolds number suppressing boundary layer transition. Friction drag may also be
range operations the test minimum width-based Reynolds number reduced in a synergistic design process where flow control systems are
shall be greater than 3 million. used to improve fairing performance allowing for the minimization of
fairing surface area that further reduce friction drag. Pursuing these
Overarching Reynolds number sensitivity criteria related to fuel advanced flow management concepts requires an understanding of the
economy or fuel use testing were identified; 1) test conditions match boundary layer flows present on a commercial vehicle and any
the operational conditions of interest and 2) the vehicle operational variations in friction drag and limitations associated with sub-scale
Reynolds number state is known prior to the test. The second criteria wind tunnel testing. Active and passive flow control technologies may
was highlighted as critical when interpreting test results that may be used to enhance the effectiveness, alter the effective shape and
exhibit large run-to-run and segment-to-segment variability resulting reduce the wetted area of current and future fairings. However, flow
from an inability to maintain Reynolds number similarity during the control technologies cannot and should not replace a fairing. These
test. A specific area of concern is the use of oval or circular test technologies can be used to control flow separation at off-design
tracks. Data was presented that showed track testing can introduce an points such as moderate and high yaw conditions or low speed
effective change (ie misalignment) in vehicle geometry that is not operation where laminar boundary layers may dominate. Flow control
corrected for in either fuel use test method and is not representative systems can also be used to manage cooling flow in the presence of
of typical on-highway conditions. A discussion of the influence of atmospheric and vehicle interference. Interference flows were
subcritical and transitional Reynolds number state on test data was discussed in the context of platooning. Time-dependent interfering
discussed as well as the influence of wind allowances in the J1321 flows will vary in magnitude and direction with a time scale that may
and J1526 test protocols. be proportional to the vehicle speed. This time dependent flow
variability may disrupt the stabilization of the vehicle boundary layer
Reynolds number investigations related to commercial vehicle drag and flow separation patterns introducing additional time dependencies
reduction began with the NASA studies in the 1970s. By 1979 SAE into the flow on the surface of the vehicle. These effects are especially
published Recommended Practice J1252 entitled “Wind Tunnel Test problematic for a commercial vehicle that operates in a narrow
Procedure for Trucks and Buses”. This document provided the first window of economic viability. Additionally these aerodynamic
set of wind tunnel testing criteria for commercial vehicles and interference flows may mask projected fuel savings benefits that are
established the original set of Reynolds number guidelines and based on controlled aerodynamic and/or fuel consumption tests.
proposed a minimum test Reynolds number of 0.7 million. In 1981 Further investigations of aerodynamics and/or fuel consumption
Gilhaus proposed 1.6 million, Cooper proposed 2.0 million in 1982 impacted by atmospheric, traffic and platoon interference requires
and 1984, Drollinger suggested 2.0 million in 1987, Olson offered 1.7 significantly greater resources than those for a single vehicle. Flow
million in 1992 and in 1998 Hucho suggested 2.5 million. similarity requirement eliminates the wind tunnel as a viable tool and
Contributing to the body of evidence supporting an increase in the imposes a significant burden on both computational simulations and
test Reynolds number were several contributions in the 2000 to 2010 operational testing. For example, computational simulations may
period. In 2004 Storms' data suggested 3.0 million and Hjelm's paper require new boundary layer models to capture changes in local
showed 4.0 million. In 2012 the 0.7 million Reynolds number was Reynolds number and boundary layer states and operational testing
removed from J1252 and replaced with a detailed test procedure to will require extensive use of instrumentation to capture aerodynamic
determine the minimum allowable Reynolds number for the specific data to support assessment of variations in local Reynolds number and
test being conducted. The intent of the test Reynolds number boundary layer state.
assessment procedure is to allow adjustments to the minimum test
Reynolds number based on differences in the test vehicle or the
specific operational Reynolds number range of the reference
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 661

Critical to the success of each topic area is the continued boundary layers completely attached over the external surface and
development of aerodynamic test and analysis tools capable of with a thin steady wake containing vorticity. The vehicle viscous drag
capturing the Reynolds number sensitivities present with future is greater than its pressure drag.
designs. Of particular importance is the maturation of computational Aerodynamic Device - A structure or system of a vehicle for altering
simulation tools with improved friction modeling capability the aerodynamic forces acting on the vehicle.
specifically developed for commercial ground vehicles.
Accuracy - The degree to which the result of a measurement,
calculation, or specification conforms to the correct value or a
CONTACT INFORMATION standard.
Richard Wood Attached Flow - A condition in which fluid flow has not become
SOLUS-Solutions and Technologies, LLC detached (separated) from the surface of the object.
754 Suffolk Lane Base - The rearmost aft-facing vertically-aligned surface of a bluff
Virginia Beach, VA 23452 body ground vehicle.
rick@solusinc.com Base Area - Orthorgraphic projection of the ground vehicle base onto
757-486-3570 a plane perpendicular to the centerline of the vehicle.
Bias - Is the difference between the average value of the large series
ACKNOWLEDGMENT of measurements and the accepted true value.
I would like to extend my appreciation to the Buckendale Lecture Bluff Body - A non-lifting vehicle with significant amount of
Committee for giving me the honor of being the 2015 Buckendale separation of the boundary layer and with trailing wakes having
Award recipient. The committee's encouragement and support have significant lateral dimensions and normally unsteady velocity fields.
been instrumental in the development of this paper. Special thanks The vehicle pressure drag is greater than the viscous drag.
are due to Steven Wesolowski for chairing the Committee and Boundary Layer - Thin layer of fluid in the immediate vicinity of a
providing guidance and assistance in managing the paper surface where the effects of viscosity are significant and the velocity
development process. I am grateful to all members of the Committee. changes from zero at the surface to the free stream value away from
The comments and contributions by Vern Caron, Richard Hanowski, the surface.
Chris Keeney, Steven Wesolowski, and Daniel Williams are Boattail / Boat-Tail - Back end of a vehicle, which is chamfered
especially noteworthy. from both sides to form a sharp end, like it is common for the rear
part of boats.
The valuable suggestions by experts namely Rick Mihelic, Ilhan
Calibration - Process of adjusting modeling parameters in one
Bayraktar, and Maximilian Hombsch have been greatly improved the
process for the purpose of improving agreement with another process.
document. I would like to make special note of the guidance and
insight provided by Rick Mihelic during the paper development Coastdown / Coast-Down testing - Coastdown testing refers to a
process. Rick Mihelic, Naethan Eagles, Kevin Horrigan and Fred method of determining a vehicle's total resistance during an on-road
Ross are recognized for many fruitful discussions and debates on the driving maneuver (drag, rolling resistance and driveline losses).
Reynolds number topic over the past few years. Starting from high speed (greater than 60mph), the vehicle is set to
neutral and is allowed to coast until its speed drops below a certain
I special thank you to Kacy Weaver and Brandie Schandelmeier of threshhold (less than 20mph).
SAE for their critical role in facilitating the monthly conference calls Confidence - Probability that a numerical estimate will lie within a
and processing the document through MyTechZone. specified range.
Cross Flow - Wind velocity that is at an angle to the vehicle
Finally, I am most indebted to my wife Sandi and daughter Melissa, centerline.
for allowing me to “act like an engineer” during the summer of 2015.
Drag - A force that opposes the forward motion of the vehicle.
Their love, support and understanding have been my motivation.
The total force is composed of a pressure drag and a viscous drag
component.
DEFINITIONS Drag coefficient - Dimensionless drag, obtained by dividing the drag
This section contains definitions of terms and symbols used in the by the dynamic pressure and a reference area, usually the largest
document as they relate to the classes of ground vehicles covered cross sectional area of a body in fluid flow.
under this recommended practice and that operate at velocities less Dynamic Pressure - Pressure that can be obtained when decelerating
than 100 mph. These definitions may differ from those used for a flow, calculated as free stream density times velocity squared
aerospace vehicles. divided by two.
Error - The difference between a measurement and the true value of
Aerodynamics - A branch of dynamics that deals with the motion of the measurand (the quantity being measured)
air and other gaseous fluids and with the forces acting on bodies in
Experiment - Observation and measurement of a physical
motion relative to such fluids
component or system.
Aerodynamic (Streamline) Body - A nonlifting vehicle with thin
Flow Separation - A condition in which fluid flow becomes detached
662 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

from the surface of the object and takes the forms of eddies and Strouhal Number - A dimensionless number describing oscillating
vortices. flow mechanisms
Froude Number - A dimensionless parameter measuring of the ratio Systematic error - Tends to shift all measurements in a systematic
of the inertia force on an element of fluid to the weight of the fluid way so that in the course of a number of measurements the mean
element - the inertial force divided by gravitational force. The Froude value is constantly displaced or varies in a predictable way.
number is commonly used in naval architecture as it represents the Tractor - A vehicle designed primarily to pull a semi-trailer by the
ratio of flow velocity to wave propagation velocity in open channels. use of the fifth wheel that is mounted over its drive axel(s). May be
Hatchback / Hatch-Back - Back end of a vehicle, which is inclined called a truck/highway tractor to differentiate it from a farm tractor.
with respect to the ground in an angle other than 90 degree, i.e. a Trailer - A freight carrying unpowered unit pulled by a powered unit.
chamfered back end instead of a vertical vehicle base.
Truck - A vehicle which carries cargo in a body (van, tank, etc.)
Lift Force - The aerodynamic force that acts perpendicular in a which is mounted to a chassis, possibly in addition to a trailer which
vertical direction to the forward motion of the vehicle. is towed by the vehicle.
Laminar Boundary Layer - Thin layer of fluid in the immediate Truck-Trailer - A truck-trailer combination consists of a truck that
vicinity of a surface where the flow is organized in layers and any holds cargo in its body which is connected to its chassis, and which
exchange of mass or momentum takes place only between adjacent tows a trailer.
layers on a microscopic scale.
Turbulent Boundary Layer - Thin layer of fluid in the immediate
Laminar Bubble - A separation of the laminar boundary layer that is vicinity of a surface where the Reynolds stresses are much larger
constrained by the downstream reattachment of a turbulent boundary than the viscous stresses and where the flow is chaotic and there is
layer. exchange of mass or momentum across multiple layers of flow.
Laminar Flow - Fluid flow in which the fluid travels smoothly or in Turbulent Flow - A flow regime characterized by chaotic and
regular paths. The flow properties at each point in the fluid remain stochastic property changes.
constant.
Uncertainty - A potential deficiency in any phase or activity of
Mach number - A dimensionless number defining the ratio of flow the modeling or experimentation process that is due to inherent
or vehicle speed to the speed of sound. variability (irreducible uncertainty) or lack of knowledge (reducible
Model - Either a conceptual/mathematical/numerical description of uncertainty). Uncertainty is the component of a reported value that
a specific physical scenario, including geometrical, material, initial, characterizes the range of values within which the true value is
and boundary data; or a sub-scale representation of a vehicle for use asserted to lie.
in wind tunnel tests. Validation - The process of determining the degree to which a model/
Precision - Is the closeness of agreement between independent test and associated data represents the intended use.
measurements of a quantity under the same conditions. Verification - The process of establishing the accuracy of a model/
Pressure Drag - The summation of the forces acting normal to the test and associated data.
vehicle surface that oppose the forward motion of the vehicle. Viscous Drag - The summation of the forces acting tangential to the
Random Error - Is a component of the total error that, in the course vehicle surface that oppose the forward motion of the vehicle.
of a number of measurements, varies in an unpredictable way. Vortex - A spinning, often turbulent, flow of fluid, often characterized
Reference Area - Orthorgraphic projection of the vehicle maximum by spiral streamlines.
cross sectional area onto a plane perpendicular to the centerline of the Yaw Angle - The effective wind angle experienced by the vehicle
vehicle. based upon vehicle ground velocity, wind velocity and wind angle,
Repeatability - Is the precision determined under conditions where relative to the vehicle heading. Vehicle drag increases exponentially
the same operator to make measurements on identical specimens uses with increasing yaw angle.
the same methods and equipment.
Reverse Flow - Is typically a mechanism of boundary layer
NOMENCLATURE
separation and occurs when the portion of the boundary layer closest
to the wall reverses in flow direction. a - acceleration

Reynolds Number - A dimensionless number that gives a measure A - reference area


of the ratio of inertial forces to viscous forces and consequently AB - base area
quantifies the relative importance of these two types of forces for AS - surface area
given flow conditions.
BL - boundary layer
Semitrailer - Truck trailer supported at the rear by its own wheels
c - speed of sound, Standard property of dry air at 59 °F (15 °C) =
and at the front by a fifth wheel mounted to a tractor or dolly.
761 mph (340.29 m/s)
Stagnation Point / Line - A location on the surface where the fluid is
CD - coastdown
brought to rest (zero velocity).
CD - coefficient of drag, D / (Q A)
Separation Point / Line - A location where the portion of the
boundary layer closest to the wall reverses in flow direction. CD,ref - coefficient of drag, reference value
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 663

Cf - skin friction coefficient s - second


CF - coefficient of a force, F / (Q A) St - Strouhal number, (ωL) / U
CM - coefficient of a moment, M / (Q A L) Str - Slope of transcritical Reynolds number angle, ΔCD / (ΔRe / 106)
Cp - coefficient of pressure t - time
ΔCD - change in coefficient of drag T - temperature
ΔCD,avg - change in wind averaged coefficient of drag u - local velocity
ΔRe - change in Reynolds number u* - friction velocity
D - drag force U - free stream velocity
ft - foot w - vehicle width
F - force WT - wind tunnel
g - acceleration due to gravity x - longitudinal distance from forward most point of body
k - roughness height y - distance normal to body
kg - kilogram δ - boundary layer thickness
l - component reference length δ* - boundary layer displacement thickness
L - vehicle reference length or length in general ε - roughness height
m - mass η - non-dimensional edge radius, r / A0.5
M - Moment μ - dynamic viscosity
Ma - Mach number, Ma = U∞ / c ν - kinematic viscosity, dynamic viscosity divided by density,
mph - miles per hour standard property of dry air at 59 °F (15 °C) = 1.572 × 10−4 ft2/s
(14.55 × 10−6 m2/s)
p - pressure
q - momentum thickness
Q - dynamic pressure, (0.5ρU2)
ρ - density, standard property of dry air at 59 °F (15 °C) = 23.77 ×
r - radius of curvature
10−4 slug/ft3 (1.225 kg/m3)
Re - Reynolds Number, (U∞L) / ν∞
τ - shear stress
ReA - Reynolds Number based on A0.5
ω - frequency of vortex shedding
Rex - Reynolds Number based on length (x)
Reblt,x - Boundary layer transition Reynolds number; based on length SUBSCRIPTS
(x) ∞ - freestream
Recr,x - Critical Reynolds number; based on length (x) B - base
Rek - Roughness Reynolds number; based on roughness height and D - diameter
velocity at the top of the roughness, (uk k)/ νk
F - forebody
Rel - Reynolds number based on component length (l)
FS - full scale
ReL - Reynolds number based on model/vehicle length (L)
l - component reference length
Rer - Reynolds number based on local radius (r)
lam - laminar
Ret - Transition Reynolds number
tot - total
Retr,x - Transcritical Reynolds number; based on length (x)
turb - turbulent
Rew - Reynolds number based on model/vehicle width (w)
x - longitudinal distance from forward most point of body
Req - Reynolds number based on momentum thickness (q)
Rr - Roughness ratio, ε/L
664 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

APPENDIX
APPENDIX - WIND TUNNELS
A listing of wind tunnels can be found in the documents contained in table A.1. The majority of automobile manufacturers have wind tunnels capable
of supporting commercial vehicle aerodynamic testing. Selected wind tunnels that may be used to support commercial vehicle aerodynamic studies
are contained in tables A.2, A.3, A.4, A.5, A.6. All tunnels listed in this appendix are subsonic facilities that have test-section widths and heights
greater than 1.0 meter.

Table A.1. Documents containing a listing of wind tunnels.

Table A.2. Asian Wind Tunnels

Table A.3. Canada, Israel Russia and South America Wind Tunnels
Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 665

Table A.4. European Wind Tunnels


666 Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015)

Table A.5. United Kingdom Wind Tunnels


Wood / SAE Int. J. Commer. Veh. / Volume 8, Issue 2 (October 2015) 667

Table A.6. United States Wind Tunnels

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or
otherwise, without the prior written permission of SAE International.

Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE International. The author is solely responsible for the content of the paper.
LECTURES
1954 KENNETH W. GORDON SP-130 “Design Evaluation & 1982 RAY W. MURPHY SP-506 “Endurance Testing of Heavy
Selection of Heavy-Duty Rear Axles” Duty Vehicles”
1955 No Lecture 1983 HANS J. BAJARIA SP-533 “Integration of Reliability,
Maintainability and Quality Parameters in Design”
1956 CARL GEORGE ARTHUR ROSEN SP-131 “The Role of the
Turbine in Future Vehicle Powerplants” 1984 JAMES D. SYMONS SP-563 “Dynamic Sealing Systems for
Commercial Vehicles”
1956 WILLIAM PEARSE MICHELL SP-132 “New Drive Lines for
New Engines” 1985 THOMAS D. GILLESPIE SP-607 “Heavy Truck Ride”
1957 ROBERT M. RIBLET and CHARLES M. KITSON SP-133 1986 TREVOR O. JONES SP-647 “Commercial Vehicle
“Bearing Application for Heavy-Duty Axles” Electronics”
1958 No Lecture 1987 RICHARD DROLLINGER SP-688 “Heavy Duty Truck
Aerodynamics”
1959 OLIVER K. KELLEY SP-134 “Planetary Gearing – Basic
Design Information & Typical Application to Commercial 1988 FRED S. CHARLES and THOMAS L. FORD SP-729 “Heavy
& Military Ground Vehicles” Duty Truck Tire Engineering”
1960 GEORGE J. HUEBNER, JR. SP-172 “Computer-Based 1989 SIDNEY F. WILLIAMS, JR. and WILLIAM A. LEASURE, JR.
Selection of Balanced-Life Automotive Gears” SP-789 “Antilock Systems for Air-Braked Vehicles”
1961 V. J. JANDASEK SP-186 “The Design of Single-Stage, 1990 DANIEL J. BOSCH and JOHN D. REAL SP-824 “Heavy
Three Element Torque Converter” Truck Cooling Systems”
1962 WELLS COLEMAN SP-221 “Design and Manufacture of 1991 CHARLES R. JONES SP-868 “Heavy Duty Drivetrains - The
Spiral Bevel and Hypoid Gears for Heavy-Duty Drive System and Component Application”
Axles” 1992 WILLIAM R. CAREY SP-913 “Tools for Today’s Engineer -
1963 G. ROBERT HARTING SP-239 “Design and Application of Strategy for Achieving Engineering Excellence”
Heavy-Duty Clutches” 1993 DAVID CEBON SP-951 “Interaction between Heavy
1964 G. P. MATHEWS SP-251 “Art and Science of Braking Vehicles and Roads”
Heavy-Duty Vehicles” 1994 DAVID F. MERRION SP-1011 “Diesel Engine Design for the
1965 PHILIP J. MAZZIOTTI SP-262 “Dynamic Characteristics of 1990’s”
Truck Drive Line Systems” 1995 WESLEY M. DICK SP-1063 “All Wheel and Four Wheel
1966 WILLIAM J. SIDELKO SP-276 “An Objective Approach to Drive Vehicle Systems”
Highway Truck Frame Design” 1996 No Lecture
1967 PAUL K. BEATENBOUGH SP-284 “Engine Cooling 1997 FARHANG ASLANI, CHING-HUNG CHUANG, SHABBIR
Systems for Motor Trucks” DOHADWALA, JEFF HUANG, BIJAN KHATIB-SHAHIDI,
1968 V. G. RAVIOLO SP-341 “Planning Product” PATRICK J. LEE, DAVID S. ROHWEDER, RANDAL H.
VISINTAINER and DAVID E. WATTS SP-1310 “CAE Methods
1969 J. A. DAVISSON SP-344 “Design and Application of
and Their Application to Truck Design”
Commercial Type Tires”
1998 LEONARD C. BUCKMAN SP-1405 “Commercial Vehicle
1970 GARY L. SMITH SP-355 “Commercial Vehicle
Braking Systems: Air Brakes, ABS and Beyond”
Performance and Fuel Economy”
1999 VALERIE A. NELSON, MARY L. RANGER and PETER
1971 CHARLES M. PERKINS SP-363 “Principles and Design of
KANEFSKY SP-1541 “A Systems Approach to Engine
Mechanical Truck Transmission”
Cooling Design”
1972 PETER R. KYROPOULOS SP-367 “Human Factors
2000 RONALD P. ZIEBELL SP-1567 “Commercial Use of Military
Methodology in the Design of the Driver’s Workspace in
Truck Technology”
Trucks”
2001 VERN ANDREW CARON SP-1650 “Commercial Vehicle
1973 JOHN W. DURSTINE SP-374 “The Truck Steering System
Electronics Design”
– From Hand Wheel to Road Wheel”
2002 LEE R. ARMSTRONG SP-1727 “Electronic System
1974 RICHARD L. STAADT SP-386 “Truck Noise Control”
Integration”
1975 PHILLIP S. MYERS SP-391 “The Diesel Engine for Truck
2003 MARK G. THOMAS SP-1816 “Electronic Systems Testing
Application”
and Validation for Commercial Vehicles”
1976 ERNEST R. STERNBERG SP-402 “Heavy-Duty Truck
2004 WILLIAM J. CHARMLEY 2004-01-2708 “The Federal
Suspensions”
Government’s Role in Reducing Heavy-Duty Diesel Engine
1977 RAYMOND E. HELLER SP-413 “Truck Electrical Systems” Emissions”
1978 KENNETH W. CUFFE SP-425 “Air Conditioning & Heating 2005 STEPHEN J. CHARLTON 2005-01-3628 “Developing Diesel
Systems for Trucks” Engines to Meet Ultra-low Emission Standards”
1979 MARTIN J. HERMANNS SP-437 “Front Drive Systems for 2006 MATTHEW BAUS, ANTHONY COOK, and DAVID SCHALLER
4WD Light Trucks” 2006-01-3545 “Integrating New Emissions Engines into
1980 JOHN R. KINSTLER SP-454 “Wheels for Commercial Commercial Vehicles: Emissions, Performance &
Vehicles” Affordability”

1981 JOHN C. WALTER SP-479 “A Guide for Powerplant 2007 DEBORAH FREUND 2007-01-4298 “Foundations of
Installation in Trucks” Commercial Vehicle Safety: Laws, Regulations and
Standards”
LECTURES CONTINUED
2008 PETER L. GODDARD 2008-01-2680 “System Safety Applied
to Vehicle Design”
2009 ALI F. MALEKI 2009-01-2924 “Embedded Software
Engineering in Automotive and Truck Electronics”
2010 MARK P. ZACHOS 2010-01-2053 “Merge Ahead: Integrating
Heavy-Duty Vehicle Networks with Wide Area Network
Services”
2011 RICHARD JOSEPH HANOWSKI 2011-01-2305 “The
Naturalistic Study of Distracted Driving: Moving from
Research to Practice”
2012 DANIEL EUGENE WILLIAMS SP-2337 “Multi-Axle Vehicle
Dynamics”
2013 DONALD WAYNE STANTON 2013-01-2421 “Systematic
Development of Highly Efficient and Clean Engines to Meet
Future Commercial Vehicle GHG Regulations”
2014 MEHDI AHMADIAN 2014-01-2408 “Integrating
Electromechanical Systems in Commercial Vehicles for 
Improved Handling, Stability, and Comfort”
2015 RICHARD WOOD 2015-01-2859 “Reynolds Number Impact
on Commercial Vehicle Aerodynamics and Performance”

P151240

You might also like