Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/357972801

A Comparative Study of Tail Air-Deflector Designs on Aerodynamic Drag


Reduction of Medium-Duty Trucks

Preprint · January 2022

CITATIONS READS

0 2,304

3 authors:

Wei Gao Deng Zhaowen


Hubei University of Automotive Technology Hubei University of Automotive Technology
22 PUBLICATIONS 44 CITATIONS 23 PUBLICATIONS 34 CITATIONS

SEE PROFILE SEE PROFILE

Yuping He
Ontario Tech University
181 PUBLICATIONS 1,821 CITATIONS

SEE PROFILE

All content following this page was uploaded by Yuping He on 20 January 2022.

The user has requested enhancement of the downloaded file.


A Comparative Study of Tail Air-Deflector Designs on Aerodynamic Drag
Reduction of Medium-Duty Trucks
Wei Gao1, Zhaowen Deng1, Yuping He2*
1Collegeof Automotive Engineering, Hubei University of Automotive Technology, Shiyan, China
2Department of Automotive and Mechatronics Engineering, University of Ontario Institute of

Technology, Oshawa, Canada


*e-mail: yuping.he@uoit.ca
Abstract: With the rapid development of highway networks and logistics industry in China, medium-
duty trucks have been increasingly used for highway and urban area freight transportation. The
aerodynamic drag reduction of medium-duty trucks is of significant importance for improving highway
transportation efficiency, enhancing fuel economy, and reducing greenhouse emissions. This paper
proposes, compares and evaluates seven designs of tail air-deflector, which are devised to reduce
aerodynamic drag of a medium-duty truck. The impacts of the shapes and/or configurations of tail air-
deflector on the aerodynamic characteristics of the truck are studied using computational fluid dynamics
(CFD) simulation. The results attained from the comparative study will provide guidelines for the design
of airflow control devices for commercial vehicles to effectively reduce aerodynamic drag.
Keywords: medium-duty trucks; tail air-deflectors; aerodynamic drag reduction; comparative study;
parametric study; computational fluid dynamics; modelling and simulation
1. Introduction
Articulated heavy vehicles (AHVs) and heavy-duty trucks are commonly used to transport freight between
major cities via highways. The popularity of these large commercial vehicles is mainly attributed to their
flexibility and transport efficiency. However, these large vehicles lack the maneuverability to travel on
minor roads or in urban areas with narrow streets and tight corners. In urban areas, smaller vehicles need
to be used for deliveries, thereby resulting in double handling of freight and reducing the efficiency of the
over-all transport system [1-2].
The past two decades witnessed the rapid development of highway networks and logistics industry in
China [3]. With the present depletion of non-renewable carbon-based fuels and increasing pollution
concerns, it is of foremost importance to improve the fuel economy of road vehicles [4]. Aerodynamics
and, in particular, aerodynamic drag plays a key role in enhancing the fuel economy of vehicles operating
on highways [5-6]. Medium-duty trucks have been increasingly used for highway freight transportation
[3]. These medium-duty trucks exhibit good maneuverability and effectively operate in urban areas
without using smaller vehicles for goods deliveries. This will improve overall transportation efficiency
without double handling of freight [2].
The aerodynamic drag reduction of medium-duty trucks is of importance for improving highway
transportation efficiency, enhancing fuel economy, and reducing greenhouse emissions. However,
medium-duty trucks exhibit worse fuel economy at high speeds than other road vehicles, such as
passenger cars, mainly due to their poor streamlined body designs. It is reported that fully loaded
commercial vehicles traveling at 100km/h consume 65% of total fuel to overcome aerodynamic drag [7-
8]. It has been reported that the average annual mileage of commercial vehicles varies between 130000
and 160000 km [9]. Therefore, for commercial medium-duty trucks, any slight reduction of aerodynamic
drag will lead to considerable fuel savings.
As shown in Figure1, the aerodynamic drag on the AHV mainly comes from the following four areas: 1)
the forebody stagnation region, 2) the gap region between the tractor and the trailer, 3) the under-body
1
region of the truck and the trailer, and 4) the rear end of the trailer [10]. Various aerodynamic
supplementary devices have been proposed to reduce the drag of AHVs. As shown in Figure 2, these add-
on devices include fairings, side skirts, roof tapering, edge rounding, base flaps, etc. [11]. The effect of the
above drag reduction devices on AHVs has been investigated. Mazyan studied the aerodynamic
performance of an AHV with front and rear drag reduction devices by numerical simulation [12]. The
simulation showed a drag reduction of about 21%. To evaluate the drag reduction property of side skirts,
two different types of the device were added on two 1/8 scale heavy-duty truck models [13]. The drag
coefficients of the two trucks were reduced by 5.3% and 4.7%, respectively. Cab-roof fairings (CRFs) have
been developed to reduce the drag of heavy truck forebodies [14]. The drag coefficient reduction rates
are approximately 20% and 23% for 15-ton and 5-ton heavy vehicle models installed with the advanced
bio-inspired CRF, respectively. Moreover, the driving stability of the heavy vehicles was enhanced by
adding the CRFs. Khosravi et al. investigated the influences of supplementary devices, such as deflector,
cab and back vanes, front fairings etc., on the aerodynamic characteristics of AHVs [15]. The results
showed that the drag coefficient of the AHV was reduced by 41% as all the add-on devices were installed
at their best positions. A modified boat tail with a lower inclined air deflector was proposed to reduce the
drag of heavy vehicles and obtained about 9.02% drag reduction at maximum [16].

Figure 1. Aerodynamic drag contributing regions on an AHV [10].

Figure 2. Aerodynamic add-on devices on trucks [11].

2
From the above literature review of the state-of-the-art studies on the aerodynamic drag reduction of
commercial vehicles, it is found that studies on the effects of various add-on devices on drag reduction
for AHVs and heavy-duty trucks have been carried out using numerical and experimental methods.
Moreover, several drag reduction devices, such as cab-roof fairings, side skirts, etc., have been applied
to AHVs and heavy-duty trucks [10, 17]. Heavy-duty trucks, e.g., tractor/semi-trailer combinations, exhibit
unique aerodynamic characteristics. Due to the multiple vehicle unit configurations, there exists a relative
yaw angle between the tractor and the trailer, which may lead to a significant increase of drag coefficient
[17]. The available research results and airflow control devices for heavy-duty trucks can’t be applied to
medium-duty trucks directly. For example, the optimum flap angle of the base flap design is sensitive to
Reynolds number and truck geometry [8]. Therefore, with the similar base flap shape design, the desired
parameter selection for a medium-duty truck may be different from the counter part for a heavy -duty
truck. Of course, the existing research results and design methods of airflow control devices for heavy-
duty trucks provide valuable guidelines for the development of these devices for medium-duty trucks.
There exist various gaps between heavy- and medium-duty trucks. In addition to total mass, the
differences between heavy- and medium-duty trucks are detailed shapes and total dimensions, which
pose significant impacts on the aerodynamic characteristics of these vehicles [18]. As aforementioned,
medium-duty trucks are increasingly used for highway freight transportation between major population
centers in China. Moreover, as mentioned above, various air-deflectors have been developed to decrease
the drag of heavy-duty trucks. Cab fairing have been extensively and deeply investigated and applied to
various kinds of heavy-duty trucks. As shown in Figure 1, for tractor/semi-trailer combinations, the rear
end of the trailer is one of the main areas that generates aerodynamic drag [10]. However, few attempts
have been made to study the effect of trail air-deflectors on aerodynamic drag reduction of medium-
duty trucks. All the aforementioned facts motivated this current research.
In the tail air-deflector design, the following fundamental factors should be considered:
shapes/configurations, width, and installation angle. This paper proposes, evaluates and compares seven
designs of tail air-deflector, which are devised to reduce aerodynamic drag of a medium-duty truck. We
investigate the drag reduction effects of shape or configuration of tail deflectors, flap width, and flap
installation. An attempt is made to identify the most effective and feasible aerodynamic drag reduction
methods and devices for medium-duty trucks.
The rest of the paper is organized as follows. In section 2, using a CFD method, we generate effective
models to represent the medium-duty truck, and design seven different shapes or configurations of tail
air-deflector. Section 3 compares the effects of the seven air-deflectors in relation with a baseline design
on aerodynamic drag reduction of the truck, and identifies the most effective air-deflector design. In
section 4, for the most effective air-deflector design, a sensitivity analysis is carried out considering two
structural parameters of the air-deflector, i.e., flap width and top/bottom flap installation angle. The
conclusions are drawn in section 5.

2. Modelling of the Medium-Duty Truck with and without Tail Air-Deflector


CFD-based vehicle models and simulations have been increasingly applied to the design and evaluation
of aerodynamic supplementary devices in a virtual environment prior to a physical full-size prototype
being fabricated and wind-tunnel tests [20-24]. To cost-effectively evaluate the aerodynamic
performance of the proposed tail air-deflectors, three-dimension (3D) CAD (computer aided design)
medium-duty truck models are generated and numerical simulations are conducted. The modelling

3
process consists of the following essential steps: 1) generating 3D CAD models, 2) developing a virtual
wind tunnel (computation domain), and 3) settings of boundary conditions and solver parameters.
2.1 3D Medium-Duty Truck Models
The dimensions of the medium-duty truck are set to 5500 mm in length, 1600 mm in width, and 2000
mm in height. Considering the limitation of computer hardware resources and the trade-off between the
computational efficiency and simulation accuracy, a 1/10 scale medium-duty truck model is developed
and studied in this research. In the modelling, we ignore the effects of door handles, rearview mirrors,
suspensions, braking system, exhaust system, and other components. The 3D models are assumed to be
symmetrical with respect to the longitudinal central plane of the vehicle. To reduce simulation time, half-
vehicle models are generated using CATIA software. Figure 3 shows the eight half-vehicle models,
including a baseline model without tail air-deflector and those with sever different shapes or
configurations of trail air-deflectors. The seven tail air-deflectors are selected considering the
operational and economic benefits and the applicability of these devices. Some of the original designs of
the associated shapes/configurations come from the research and development of medium-duty trucks
of JAC Motors [25]. The dimensions of the deflectors are provided in Table 1.
Table 1. Shapes/configurations and dimensions of tail air-deflectors.
Shapes/configurati
Dimensions of deflectors (scale ratio 1:10)
ons of deflector
The angle between the deflector and the top surface of the trailer is 10°, 70mm in length,
Rectangle
40mm in width, and 5mm in thickness
The angle between the deflector and the top surface of the trailer is 10°, 70mm in major
Ellipse
axis length, 40mm in minor axis length, and 5mm in thickness
The angle between the deflector and the top surface of the trailer is 10°, 70mm, and 40mm
Right triangle
in side lengths, and 5mm in thickness
The angle between the upper deflector and the top surface and the lower deflector and
Top and bottom
the bottom surface of the trailer are all 10°, 70mm in length, 40mm in width, and 5mm in
rectangle
thickness
The angle between the upper deflector and the top surface and the lower deflector and
Base flaps the bottom surface of the trailer are all 10°, the angle of the side flap with the side surface
of the trailer is 0°, 70mm in length, 40mm in width, and 5mm in thickness.
The bottom radius is 81.25mm, 70mm in height. The distance between the cylinder side
Partial cylinder
and the rear of the vehicle is 40mm.
The bottom radius is 81.25mm, 70mm in height. The distance between the cylinder side
Cylinder truncated
and the rear of the vehicle is 40mm. The Angle between the cutting plane and the
body
horizontal plane is 30°

4
(a) (b)

(c) (d)

(e) (f)

(g) (h)

Figure 3. Half-vehicle CAD model with/without tail air-deflector: (a) baseline, (b) rectangle, (c) ellipse, (d)
triangle, (e) top and bottom rectangle, (f) base flaps, (g) partial cylinder, (h) cylinder truncated body.

2.2 Virtual Wind Tunnel


The CAD models of the medium-duty truck with and without tail air-deflectors are each fixed in a virtual
wind tunnel, which is actually a 3D computational domain created in ANSYS Workbench® (ANSYS Inc.,
2020). The virtual wind tunnel is a cuboid that contains a medium-duty truck model as shown in Figure
4. The dimensions of the virtual wind tunnel are set as follows: 1) the inlet and outlet of the wind tunnel
are 3L upstream and 7L downstream of the medium-duty truck model, respectively (where L is the
overall length of the medium-duty truck model); 2) The distance between the longitudinal sidewall of
the wind tunnel and the side surface of the medium-duty truck is 4W (where W is the overall width of

5
the medium-duty truck model); and 3) The top wall of the wind tunnel is at a distance of 4H from the
top surface of the model (where H is the overall height of the medium-duty truck model).

Figure 4. Virtual wind tunnel and a medium-duty truck mode.

The computational domains of all the models are discretized with unstructured hybrid meshes including
tetrahedron and prism elements. To improve the accuracy of numerical simulation, the meshes around
the medium-duty truck models are refined. Moreover, to examine the effect of near-wall boundary of
air flow and accelerate the convergence of numerical simulation, the prismatic boundary layers are
added to the medium-duty truck body surfaces. The mesh model for the truck with the tail air-deflector
with the configuration of cylinder truncated body is shown in Figure 5.

Figure 5. Meshes around and on the medium-duty truck with the tail air-deflector with the
configuration of cylinder truncated body.
2.3 Boundary Conditions and Solver Parameter Setting
To investigate the aerodynamic characteristics of the medium-duty truck with and without tail air-
deflector, CFD simulations of proposed models are conducted in the virtual wind tunnel. The definition
of the virtual wind tunnel boundary conditions is illustrated in Figure 6. The setting of the boundaries is
listed in Table 2. The pressure-based solver and steady states are chosen for these simulations. The
realizable k −  turbulence model, standard wall function, second-order upwind solution method, and
6
effective scheme are used in the study. The residual value and the iteration number are set to 10 -6 and
2000, respectively.
Table 2. Setting of boundary conditions.
Boundary of domain Conditions
Velocity inlet: velocity magnitude V=30 m/s, turbulent intensity I=0.5%, turbulence
Inlet
Length Scale l =0.667m
Pressure outlet: gauge pressure P=0 Pa, turbulent intensity I=5%, turbulence length
Outlet
scale l =0.667m
Ground Moving wall: speed U=30m/s
Top and side Stationary wall, no-slip
Truck body Stationary wall, no-slip
Symmetry plane Stationary wall, no-slip

Figure 6. Boundary conditions of the virtual wind tunnel.


3. A Comparison of Simulation Results for Different Tail Air-Deflector Designs
To evaluate the effects of different air-deflector designs on drag reduction of the truck, we compare the
CFD simulation result of the baseline vehicle against those of the vehicle with different air-deflector
designs as shown in Figure 1. The aerodynamic drag is usually determined by the formulation [26]:
1
𝐹𝑑𝑟𝑎𝑔 = 2 𝜌𝐶𝐷 𝐴𝑓 𝑉𝑟2 (1)
where 𝜌 the air mass density, 𝐶𝑑 the drag coefficient, 𝐴𝑓 the vehicle frontal area, 𝑉𝑟 the vehicle velocity
relative to the wind. It is indicated that the drag coefficient, 𝐶𝑑 , represent the combined effects of all
contributing factors causing the aerodynamic drag [24]. Thus, the drag coefficient is used as the unified
indicator to evaluate the aerodynamic drag performance of the truck with and without air-deflector.
The drag coefficient and drag reduction for each of the designs are shown in Table 3. Note that, as shown
in the table, for the purpose of simplicity, the baseline vehicle without air-deflector to the vehicle with
the cylinder truncated body shaped air-deflector are denoted as 1 to 8. As expected, the simulation
result indicates that the drag coefficients of the truck with the seven designs of air-deflector are all
smaller than that of the baseline design. The most effective air-deflector is design 6, i.e., base flaps, with
the drag reduction of 6.71% in comparison with design 1. Among the seven designs (2 to 8), design 8 is
the most sophisticated device with the highest complexity in design. It is expected that design 8 should
7
exhibit superior performance among the seven designs. Surprisingly, among the seven designs, design 8
shows the worst drag reduction effect with the drag coefficient of 0.77691, which is reduced by only 1.1%
compared against the drag coefficient of 0.78557 of design 1.
To seek the root causes resulting in the difference in drag coefficient values of the eight designs, Figure
7 compares these designs in terms of the wake flow velocity vectors on the longitudinal symmetry plane.
For design 1, as shown in Figure 7 (a), a large longitudinal vortex is formed, and the vortex center is
located close to the rear vehicle body wall at the lower left corner of the velocity field. It appears that,
for design 1, the wake turbulent flow with a higher Reynolds number is faster than those of other designs.
Moreover, the wake turbulent flow field of design 1 seems larger than other designs. The largest drag
coefficient of design 1 may be attributed to the larger wake turbulent flow field and the faster turbulent
flow. Similarly, comparing the wake turbulent flow fields of designs 6 and 8, we note that the wake
turbulent flow and the vortex velocity field of the former are slower and smaller than the counterparts
of the latter. The observation justifies that the drag coefficient of design 6 is smaller than that of design
8.
Table 3. Drag coefficient values of the medium-duty truck with and without tail air-deflector.
Designs Shapes/configurations of tail air-deflector Drag coefficient Drag reduction
1 Baseline design without tail air-deflector 0.78557 —
2 Rectangle 0.76076 3.16%
3 Ellipse 0.77250 1.16%
4 Right triangle 0.77287 1.62%
5 Top and bottom rectangle 0.74917 4.63%
6 Base flaps 0.73286 6.71%
7 Partial cylinder 0.77115 1.84%
8 Cylinder truncated body 0.77691 1.10%

In the operation of a road vehicle, aerodynamic drag exerted on the vehicle will consume energy. For the
vehicles shown in Figure 1 and listed in Table 3, it is expected that under the same operating condition,
the diversity of aerodynamic drags for different designs is attributed to the difference of the tail air-
deflectors. The generation of wake flow results in energy consumption, which can be represented by
turbulence kinetic energy (TKE) dissipation. Thus, examining wake flow TKE distributions on the
longitudinal symmetry plane for all eight designs, we may evaluate the effects of different air-deflectors
on aerodynamic drag reduction of the truck.
Figure 8 illustrates the wake flow TKE distribution on the longitudinal symmetry plane for each of the
eight designs. A close observation of Figure 8 discloses the following facts: 1) The total area with few non-
zero TKE contoured regions for design 6, as shown in Figure 8(f), is the smallest among the eight designs;
for design 6, the contoured region in light green color represents the highest TKE distribution area, which
is ranked the lowest in terms of TKE intensity among the respective regions of the eight designs; 2) The
total area with a few non-zero TKE contoured regions for design 1, as shown in Figure 8(a), is the largest
among the eight designs; the contoured regions (in brown, yellow, light yellow, weak yellow, green, and
light green colors) denote the high TKE distribution areas, implying a high rank of design 1 in terms of TKE
intensity among the eight designs; 3) Similarly, as shown in Figure 8(h), the total area with a few non-zero
TKE contoured regions for design 8 is larger than the majority of the eight designs; among the eight

8
designs, design 8 takes a higher rank in terms of TKE intensity. The larger the total area with non-zero TKE
contoured regions becomes, and the higher the TKE intensity is, the more energy the wake flow dissipates.
Examining the wake flow TKE distribution on the longitudinal symmetry plane for the eight designs, we
can conclude: 1) the analysis of the wake flow TKE dissipation of the eight designs is consistent with the
result listed in Table 3; 2) compared with design 1 (the baseline vehicle), installing a tail air-deflector will
either reduce the wake flow area or decrease the wake flow TKE intensity, resulting in aerodynamic drag
reduction; 3) the design (i.e., the shape or configuration) of a tail air-deflector has a signification impact
on the wake flow TKE dissipation; 4) among the seven designs of tail air-deflectors, design 6 (i.e., the
base flaps) is the most effective device on the aerodynamic drag reduction of the medium-duty truck.
Note that the above conclusion is derived from the simulation results based on the theoretical basis
shown in Equation (1). It is recognized that Equation (1) is valid for the ambient conditions, which are
characterized by low turbulence and steady state. In reality, the vehicle may experience an unsteady
environment, caused by the effects of the natural wind and by traffic, which may introduce high levels
of turbulence. Under conditions of high-level turbulence, additional unsteady drag component may arise,
and further investigations need to be conducted.
The side area of the medium-duty truck will not be greatly increased due to installing the tail air-deflectors,
therefore, it will have negligible influence on the crosswind stability of the vehicle. The above benchmark
is conducted mainly considering the shape or configuration of air-deflector. It is indicated that the air-
deflector with the structure of base flaps (i.e., design 6) outperforms the other six designs in reducing
aerodynamic drag of the truck. In the next section, design 6 is further examined by means of refining
relevant structural parameters to further improve the performance of the deflector.

9
(a) (b)

(c) (d)

(e) (f)

(g) (h)

Figure 7. Wake flow velocity vectors on the longitudinal symmetry plane: (a) design 1, (b) design 2,
(c) design 3, (d) design 4, (e) design 5; (f) design 6, (g) design 7, and (h) design 8.
10
(a) (b)

(c) (d)

(e) (f)

(g) (h)

Figure 8. Wake flow TKE distribution on longitudinal symmetry plane: (a) design 1, (b) design 2, (c)
design 3, (d) design 4, (e) design 5, (f) design 6, (g) design 7, and (h) design 8.
4. Effects of Structural Parameters of Base Flaps on Drag Reduction
11
In this section, the effects of the base flaps with various flap widths and top/bottom flap installation
angles on drag reduction of the medium-duty truck are examined by means of CFD simulation. The
schematic representation of the flat width and top/bottom flap installation angle for the base flaps is
shown in Figure 9, where W and  denote the flap width and top/bottom flap installation angle,
respectively.

Figure 9. Schematic diagram showing flap width and top/bottom flap installation angle of base flaps.

4.1 Effect of Top/Bottom Flap Installation Angle


To investigate the effect of the installation angle for base flaps on drag reduction of the medium-duty
truck, the aerodynamic characteristic of the vehicle is analyzed when the installation angle varies from
0° to 30°, and the flap width remains fixed (i.e., W = 40mm for the vehicle model with the scale ratio of
1:10). The settings of the virtual wind tunnel, boundary conditions, and solvor parameters are the same
as those in section 3.
Simulation results are provided in Table 4. It is found that the maximum drag reduction of 9.12% is
achieved at the installation angle of 5°. It should be mentioned that the percentage of drag reduction at
a given installation angle listed in Table 4 is determined by calculating the relative error between the
drag coefficient of the specified design and that of the baseline vehicle without tail air-deflector. Thus,
for the purpose of comparison, the drag coefficient of the baseline vehicle without tail air-deflector is
also listed in Table 4. Figure 10 displays the variation of the drag coefficient of the vehicle equipped with
the base flaps with the installation angle varying from 0° to 30°. As shown in Figure 10, when the
installation angle varies in the range of 0°to 30°, the drag coefficient of the vehicle with the base flaps at
any installation angle is lower than that of the baseline vehicle. Within the given variation range of
installation angle, the drag coefficient first decreases, following an increase after the valley point at the
angle of 5° until the peak point at the angle of 25𝑜 , then decreases again and ends at the installation
angle of 30𝑜 .

12
Table 4. Effect of top/bottom flap installation angle of the base flaps on drag coefficient.
Top/Bottom Flap Installation angle Drag Coefficient Drag Reduction
0° 0.73551 6.37%
5° 0.71394 9.12%
10° 0.73286 6.71%
15° 0.73623 6.28%
20° 0.75899 3.38%
25° 0.76663 2.41%
30° 0.75583 3.79%
Baseline vehicle without tail air-deflector 0.78557 —

Figure 10. Variation of drag coefficient of the vehicle with base flaps with varied top/bottom flap
installation angle.

4.2 Effect of Flap Width


To assess the effect of flap width on the aerodynamic drag reduction of the vehicle with the base flaps,
we examine the variation of the drag coefficient with respect to the varied flap width from 30mm to
50mm (for the vehicle model with the scale ratio of 1:10). It is assumed that the top/bottom flap
installation angle is fixed at 5°, which is the ‘optimal’ angle for the minimum drag coefficient with the
fixed flap width of 40 mm, as discussed in the subsection of 4.1.
Table 5 presents the simulation results in terms of the drag coefficient values with the given flap width
values. For each flap width, the aerodynamic drag reduction of the vehicle with the respective base flaps
in relation with the baseline vehicle without tail air-deflector is also provided in the table. It is observed
that, within the variation range of the flap width, the ‘optimal’ flap width is 40mm, with which the drag
coefficient attains the ‘minimum’ value of 0.71394, thereby leading to aerodynamic drag reduction by
13
9.12% compared against the baseline vehicle without air-deflector. Figure 11 visualizes the numerical
results provided in Table 5.
Table 5. Effect of flap width of base flaps on drag coefficient.

Flap Width (scale ratio 1:10) Drag Coefficient Drag Reduction


30mm 0.7416 5.60%
35mm 0.73427 6.53%
40mm 0.71394 9.12%
45mm 0.74328 5.38%
50mm 0.74043 5.75%
Baseline vehicle without tail air-deflector 0.78557 —

Figure 11. Variation of drag coefficient of the vehicle with base flaps at varied flap width.

Comparing the simulation results of design 6, i.e., the tail air-deflector with the configuration of base
flaps, provided in Tables 3, 4 and 5, we note that, on the base of ‘optimal’ flap width of 40 mm, the
‘optimal’ top/bottom flap installation angle of 5° can further reduce aerodynamic drag from 6.71% to
9.12% compared against the baseline vehicle without tail air-deflector. From the above air-deflector
shape or configuration comparison and structural parameter sensitivity analysis, we find that, in addition
to CFD simulation, designing an optimal tail air-deflector needs to conduct a comprehensive
shape/configuration topology and parameter optimization.

5. Conclusions

14
This paper presents a comparative study of tail air-deflector designs on aerodynamic drag reduction on
medium-duty trucks. Seven different designs of tail air-deflector are proposed and devised. The
comparative study is carried out by means of CFD simulations. To implement CFD simulations, eight 3D
CAD truck models with and without tail air-deflector are generated, then each of the truck model is fixed
in a virtual wind tunnel, and eventually the CFD simulations are performed after defining the boundary
conditions and setting the solver parameters. The most effective air-deflector design, i.e., the base flaps,
is identified by comparing the effects of the eight vehicle models with and without air-deflector on the
aerodynamic drag reduction of the truck. To further improve the performance of the base flaps, sensitivity
analysis is conducted considering two structural parameters, i.e., flap width and top/bottom flap
installation angle. The main findings from the research are summarized as follows:
1) Among the seven designs of tail air-deflectors, the base flaps is the most effective design, with which
the maximum drag reduction of 6.71% can be achieved in comparison with the vehicle without tail air-
deflector.
2) With the optimal values of the two structural parameters, i.e., flap width and top/bottom flap
installation angle, the aerodynamic drag of the truck with the base flap can be reduced by 9.12%.
3) Drag coefficient is an effective indicator to assess the aerodynamic characteristics of different tail air-
deflector designs. However, this indicator can’t directly provide useful information to guide improved
or modified air-deflector designs. In contrast, the analysis of turbulence kinetic energy of wake flow
may provide valuable data for air-deflector design.
4) To attain optimal tail air-deflector designs, CFD simulation-based shape/configuration topology and
parameter optimizations need to be conducted.

References
[1] Jujnovich BA and Cebon D. Path-following steering control for articulated vehicles. ASME Journal of
Dynamic Systems, Measurement, and Control, 135(3), 2013. DOI: 10.1115/1.4023396.
[2] Sanchez-Diaz I, Palacios-Argϋello L, Levandi A, et al. A time-efficiency study of medium-duty trucks
delivering in urban environments. Sustainability 2020, Vol. 12, 425; doi:10.3390/su12010425.
[3] Guo P, Hu X, Zhu Y, et al. Investigation on aerodynamic drag reduction of commercial truck based
on external styling of cab. Applied Mechanics and Materials 2013, Vol. 307, pp 186-191.
[4] Sikder T, Kapoor S and He Y. Optimizing dynamic performance of high-speed road vehicles using
aerodynamics aids. Proceedings of the ASME 2016 International Mechanical Engineering Congress
and Exposition, IECE2016, Paper No: IMECE2016-65414, V007T09A060; 9 pages, November 11-17,
2016, Phoenix, Arizona, USA.
[5] Gao W, Deng Z, Feng Y and He Y. Numerical simulation and analysis of aerodynamic characteristic
of road vehicles in platoon. Proceedings of the Canadian Society for Mechanical Engineering
International Congress 2020, CSME Congress 2020, 6 pages, June 21-24, 2020, Charlottetown, PE.
Canada.
[6] Gao W, Deng Z, Feng Y and He Y. On aerodynamic drag reduction of road vehicles in platoon. To
appear in International Journal of Vehicle Systems Modelling and Testing.
[7] McCallen R, Couch R, Hsu J, Browand F. et al. Progress in reducing aerodynamic drag for higher
efficiency of heavy duty trucks (class 7-8). SAE Technical Paper, 1999-01-2238, 1999.

15
[8] McCallen R, Salari K, Ortega J, Castellucci P. et al. DOE’s effort to reduce truck aerodynamic drag
through joint experiments and computations. SAE Technical Paper, 2005-01-3511, 2005.
[9] Cooper KR. Commercial vehicle aerodynamic drag reduction: Historical perspective as a guide. In
The aerodynamics of heavy vehicles: Trucks, buses, and trains. Lecture Notes in Applied and
Computational Mechanics, vol.19. Springer, Berlin, Heidelberg, 2004, pp.9-28.
[10] Curry T, Liberman I, Andrews LH, et al., Reducing aerodynamic drag & rolling resistance from heavy-
duty trucks: summary of available technologies & applicability to Chinese trucks. In International
Council on Clean Transportation Report. http://www.theicct.org/sites/default/ files/publications/
AERO_RR_Technologies_Whitepaper_FINAL. Accessed Oct 2012.
[11] Pourasad Y, Ghanati A, Khosravi M. Optimal design of aerodynamic force supplementary devices for
the improvement of fuel consumption and emissions. Energy & Environment, vol. 28, no.3, pp. 263-
282, 2017.
[12] Mazyan WI, Numerical simulations of drag-reducing devices for ground vehicles. Master Thesis,
American University of Sharjah,United Arab Emirates, 2013.
[13] Hwang BG, Lee S, Lee EJ, et al. Reduction of drag in heavy vehicles with two different types of
advanced side skirts. Journal of Wind Engineering and Industrial Aerodynamics, vol. 155, pp. 36-46,
2016.
[14] Kim JJ, Hong J and Lee SJ. Bio-inspired cab-roof fairing of heavy vehicles for enhancing drag reduction
and driving stability. International Journal of Mechanical Sciences, vol. 131-132, pp. 868-879, 2017.
[15] Khosravi M, Mosaddeghi F, Oveisi M, et al. Aerodynamic drag reduction of heavy vehicles using
append devices by CFD analysis. Journal of Central South University, vol. 22, no.12, pp. 4645-4652,
2015.
[16] Lee EJ and Lee SJ. Drag reduction of a heavy vehicle using a modified boat tail with lower inclined
air deflector. Journal of Visualization, vol. 20, no.4, pp. 743-752, 2017.
[17] Patten J, McAuliffe B, Mayda W, et al. Review of aerodynamic drag reduction devices for heavy
trucks and buses. NRC technical report, STT-HVC-TR-205, 11 May, 2012.
[18] Yang Yang Hu. Computational investigation of drag reduction with a rear flap on trucks. Master
Thesis, The University of Alabama at Birmingham, USA, 2012.
[19] Yang X and Ma Z. Drag reduction of a truck using append devices and optimization. Proceedings of
International Conference on parallel Computing in Fluid Dynamics, ParCFD 2013: Parallel
Computational Fluid Dynamics, pp. 332-343.
[20] Ayyagari DT and He Y. Aerodynamic analysis of an active rear split spoiler for improving lateral
stability of high-speed vehicles. International Journal of Vehicle Systems Modelling and Testing 2017,
Vol. 12, Nos. 3/4, pp. 217-239.
[21] He Y. Design of an actively controlled aerodynamic wing to increase high-speed vehicle safety. SAE
Technical Paper 2013-01-0802, 2013.
[22] Hammad M, Qureshi K and He Y. Safety and lateral dynamics improvement of a race car using active
rear wing. SAE Technical Paper 2019-01-0643.
[23] Cai J, Kapoor S, Sikder T and He Y. Effects of active aerodynamic wings on handling performance of
high-speed vehicles. SAE Technical Paper 2017-01-1592.

16
[24] Hammad M, He Y. A review of active aerodynamic control for increasing safety of high-speed road
vehicles. Proceedings of the Joint Canadian Society for Mechanical Engineering and CFD Society of
Canada International Congress 2019, London, Ontario, Canada, 6 pages.
[25] JAC Motors. Website Link (available on May 16, 2021). https://en.wikipedia.org/wiki/JAC_Motors
[26] Wong J. Theory of ground vehicles, Fourth Edition, John Wiley & sons, Inc., 2008.

17

View publication stats

You might also like