ArticlePDF Available

Abstract and Figures

Ball sports are becoming faster and more demanding than ever before, pushing traditional ball designs to their limits. In order to meet the increasing performance requirements, the ball manufacturers are producing new designs progressively. A traditional spherical football made of 32 leather panels stitched together in 1970s has now become only 14 synthetic curved panels thermally bonded (without stitches). This 14-panels ball is believed to be more spherical and performs well. The primary objectives of this study were to evaluate aerodynamic performances of footballs made of 32 panels and 14 panels. The aerodynamic forces and moments were measured experimentally for a range of wind speeds (20 km/h to 130 km/h) and the non-dimensional drag coefficient was determined and compared.
Content may be subject to copyright.
A
vailable online at www.sciencedirect.com
Procedia
Engineering
Procedia Engineering 00 (2009) 000–000
www.elsevier.com/locate/procedi
a
8
th
Conference of the International Sports Engineering Association (ISEA)
A Comparative Study of Football Aerodynamics
Firoz Alam*, Harun Chowdhury, Hazim Moria and Franz Konstantin Fuss
School of Aerospace, Mechanical and Manufacturing Engineering, RMIT University, Melbourne, VIC 3083, Australia
Received 31 January 2010; revised 7 March 2010; accepted 21 March 2010
Abstract
Ball sports are becoming faster and more demanding than ever before, pushing traditional ball designs to their limits. In order to
meet the increasing performance requirements, the ball manufacturers are producing new designs progressively. A traditional
spherical football made of 32 leather panels stitched together in 1970s has now become only 14 synthetic curved panels thermally
bonded (without stitches). This 14-panels ball is believed to be more spherical and performs well. The primary objectives of this
study were to evaluate aerodynamic performances of footballs made of 32 panels and 14 panels. The aerodynamic forces and
moments were measured experimentally for a range of wind speeds (20 km/h to 130 km/h) and the non-dimensional drag
coefficient was determined and compared.
© 2009 Published by Elsevier Ltd.
Keywords: Aerodynamics, football, wind tunnel, drag;
1. Introduction
Aerodynamics plays a prominent role in defining the flight trajectory of all high speed ball sports. Depending on
aerodynamic behaviour, the ball can be deviated from its anticipated flight path significantly resulting in a curved
and unpredictable flight trajectory. Lateral deflection in flight, commonly known as swing, swerve or curve, is well
recognized in cricket, baseball, golf, tennis, volleyball and football (soccer). In most of these sports, the lateral
deflection is produced by spinning the ball about an axis perpendicular to the line of flight. Therefore, the
aerodynamic properties of a sport ball is fundamental for the players, coaches (trainers), regulatory bodies, ball
manufacturers and even the spectators. It is no doubt that the game of football is the most popular in the world. No
other game is so much loved, played and excited spectators than football. It is played in every corner by every nation
in the world. Although, the football among all sport balls traditionally has better aerodynamic properties and
balance, however, over the years, the design of football has undergone a series of technological changes, in which
the ball has been made to be more accurate and aerodynamically efficient by utilizing new design and
manufacturing processes. Adidas, the official supplier and manufacturer of footballs to FIFA (Federation
* Corresponding author. Tel.: +61 3 99256103; fax: +61 3 99256108.
E-mail address: firoz.alam@rmit.edu.au
c
2010 Published by Elsevier Ltd.
Procedia Engineering 2 (2010) 2443–2448
www.elsevier.com/locate/procedia
1877-7058
c
2010 Published by Elsevier Ltd.
doi:10.1016/j.proeng.2010.04.013
2 F. Alam et al. / Procedia Engineering 00 (2010) 000–000
Internationale de Football Association), has applied thermal bonding to replace conventional stitching to make a
seamless surface design and an improved carcass shape by using 14 curved panels (making the ball topologically
equivalent to a truncated octahedron) instead of 32 panels previously used in the ball since 1970. It is claimed that
the ball is more spherical and performs more uniformly regardless of where it is hit. However, no independent
studies have been reported in support of this statement. Although the aerodynamic behaviours of other sports balls
have been studied by Alam et al. [1, 2], Mehta [4] and Smits et al. [5], scant information is available to the public
domain about the aerodynamic behaviour of new seamless football except experiential studies by Asai [6, 7] and
computational study by Barber et al. [8]. Therefore, the primary objective of this work is to experimentally study the
aerodynamic properties of a new seamless ball and also a traditional 32-panel ball.
Nomenclature
D Drag Force
L Lift Force
S Side Force
C
D
Drag Coefficient
C
L
Lift Coefficient
C
S
Side-Force Coefficient
Re Reynolds Number
V Velocity of Air
ρ Density of Air
A Projected Area
2. Description of Balls
Two new balls have been selected for this study. One of these two balls was made of 32 leather panels by Nike
and the other ball was made by Adidas with thermally bonded 14 synthetic panels. Both are FIFA approved balls.
The diameter of the 32-panel ball is approximately 220 mm which is inflated with three different pressures. The size
the ball is 5 (eg, the circumference of the ball is in between 27 to 28 inch and the mass is in between 14 to 16
ounce). The 32-panel ball is stitched together to provide a truncated icosahedron archimedean spherical shape. The
14-panel Adidas ball is thermally bonded machine-pressed ball without any stitches or seams, which is believed to
be more spherical compared to a 32-panel ball. The diameter of the ball is approximately 220 mm and the size of the
ball is 5. A sting mount was used to hold the ball, and the experimental set up in the wind tunnel test section is
shown in Figure 2. The aerodynamic effect of sting on the ball was measured and found to be negligible. The
distance between the bottom edge of the ball and the tunnel floor was 420 mm, which is well above the tunnel
boundary layer and considered to be out of significant ground effect.
(a) Nike made 32-panels Football (with seams and stitches) (b) Adidas made 14-panels Football (seamless)
Fig. 1. Photographs of football
2444 F. Alam et al. / Procedia Engineering 2 (2010) 2443–2448
F. Alam et al. / Procedia Engineering 00 (2010) 000–000 3
3. Experimental Set Up
In order to measure the aerodynamic properties of two footballs experimentally, the RMIT Industrial Wind
Tunnel was used. The tunnel is a closed return circuit wind tunnel with a maximum speed of approximately 150
km/h. Two mounting studs (stings) holding the ball with a six component force sensor (type JR-3) in the wind tunnel
were manufactured and purpose made computer software was used to digitise and record all 3 forces (drag, side and
lift forces) and 3 moments (yaw, pitch and roll moments) simultaneously. More details about the tunnel can be
found in Alam et al. [3]. The experimental set up of both balls in the wind tunnel is shown in Figure 2.
(a) 32-panel football (b) 14-panel football
Fig. 2. Experimental setup in the test section of RMIT Industrial Wind Tunnel
Each ball was fixed to the sting with an adhesive in order to make it very rigid. Three forces (drag, lift and side
force) and their corresponding moments were measured simultaneously under a range of speeds (20 km/h to 130
km/h within an increment of 20 km/h). The aerodynamic forces are defined as drag (D) acting in the opposite
direction to the wind, lift (L) acting perpendicular to the wind direction, and the side force acting (S) sideways based
on a frontal view. The measured aerodynamic forces were converted to non-dimensional drag coefficient (C
D
), the
lift coefficient (C
L
) and the lateral-force coefficient (C
S
), using the formula as defined in Eqs. 1 to 3.
AV
D
C
D
2
2
1
ρ
=
(1)
AV
L
C
L
2
2
1
ρ
=
(2)
AV
S
C
S
2
2
1
ρ
=
(3)
4. Results
4.1. Flow Visualization
In order to understand the flow structure around a 32-panel ball and a 14-panel seamless ball, the airflow was
visualized using smoke (see Figure 3).
F. Alam et al. / Procedia Engineering 2 (2010) 2443–2448 2445
4 F. Alam et al. / Procedia Engineering 00 (2010) 000–000
(a) Airflow around the 32-panels football (b) Airflow around the 14-panels football
Fig. 3. Airflow structure of football with smoke flow visualization
Due to the roughness created by the seams in 32-panel ball, the airflow over ball became turbulent and
subsequently generated favorable pressure gradient and delayed flow separation as shown in Figure 3(a). The
airflow appears to be separated at around 100° from horizontal direction. Generally, the flow separates at around 90º
from the horizontal for a smooth surfaced sphere. For the 14-panel seamless and stitch-less ball, the surface is more
spherical and smooth. The ball behavior is very similar to a smooth sphere. As shown in Figure 3(b), the airflow
separates at around 90º from the horizontal as in the case of a smooth sphere. Therefore, the 14-panel ball can
potentially generate more aerodynamic drag at low speeds.
4.2. Aerodynamic Drag
The aerodynamic drags for the 32-panel Nike ball under 14.5 pound per square inch (psi) air pressure, 14-panel
Adidas ball under two different air pressures (13 and 14.5 psi) and a sphere for a range of Reynolds number varied
by wind speeds are shown in Figure 4. Two different pressures were chosen to see if there was any significant effect
of pump up pressure on aerodynamic properties. There is no notable variation in drag for the Adidas ball. Both balls
have similar trend, however, a minor fluctuation of drag was noted for the 32-panel Nike ball. The Nike ball
displayed more aerodynamic drag compared to the Adidas ball in the range of 60 km/h to 120 km/h. The
aerodynamic drag for the smooth sphere has clearly demonstrated notable variation and also undergone transition
from laminar to turbulent flow. No transition for the Adidas and Nike balls was noted (see Figure 4). In contrast, the
flow transition for the sphere is clearly visible in the plot in Figure 4.
The drag coefficient C
D
for the Adidas, Nike and a sphere is shown in Figure 5. The average C
D
value for both
balls is around 0.23 at speeds above 60 km/h. The transition (laminar boundary layer to fully turbulent boundary
layer) for both balls occurs in the range of Reynolds numbers 1.1y10
5
to 3y10
5
. In contrast, the boundary layer
undergoes transition for a smooth sphere at Reynolds numbers of 2.9y10
5
to 4.6y10
5
which is notably different from
flow regime around a football.
The boundary layer transition of a football is occurred much earlier compared to a smooth sphere. The results
from this study have agreed well with the published data by Asai et al. [9]. Although, the Nike 32-panel ball
displays relatively higher C
D
between 60 to 120 km/h speeds, the variation in drag coefficients for the Adidas 14-
panel ball and Nike 32-panel ball was not significant. It is clear from Figures 4 and 5 that the C
D
for the 32-panel
ball fluctuates more compared to the C
D
value of the 14-panel Adidas ball as it is believed to be more spherical than
the Nike 32-panel ball. The small variation in pump up pressure has virtually no effect on the aerodynamic drag as
shown in Figure 5.
2446 F. Alam et al. / Procedia Engineering 2 (2010) 2443–2448
F. Alam et al. / Procedia Engineering 00 (2010) 000–000 5
Drag force as a function of wind speed
0
1
2
3
4
5
6
7
8
0 20 40 60 80 100 120 140
Speed (km/h)
Drag Force (N)
14-Panel (Adidas) @ 13 psi
14-Panel (Adidas) @ 14.5 psi
32-Panel (Nike) @ 14.5 psi
Smooth Sphere
Fig. 4. Aerodynamic drag of balls and a smooth sphere
Drag coefficient as a function of Reynolds Number
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0.45
0.5
0 100000 200000 300000 400000 500000 600000
Reynolds Number (Re)
Drag Coefficient (C
D
)
14-Panel (Adidas) @ 13 psi
14-Panel (Adidas) @ 14.5 psi
32-Panel (Nike) @ 14.5 psi
Smooth Sphere
Fig. 5. Drag coefficients of balls and a smooth sphere
F. Alam et al. / Procedia Engineering 2 (2010) 2443–2448 2447
6 F. Alam et al. / Procedia Engineering 00 (2010) 000–000
5. Discussion
The C
D
value largely depends on the roughness of the ball exterior surface and the seams. They can cause
additional drag due to the boundary-layer separation. The results indicate that the C
D
value for a football is in
between a smooth sphere and a golf ball. The golf ball data was not shown here, however for more details, see [6].
As the speeds of the football are generally in the range of 90 km/h to 130 km/h during a free kick or long shot, the
C
D
value of 32-panel or 14-panel balls are expected to be the same. However, the C
D
value can be in the transition
zone when the ball is kicked for a short pass.
6. Conclusions
The following conclusions were made from the work presented here:
The drag coefficient of a non-spinning football is approximately 0.40 at low speeds (below 30 km/h) and 0.23 at
high speeds (over 60 km/h).
The 32-panel ball has slightly higher drag at high speeds compared to the 14-panel ball.
The drag coefficient of the 32-panel ball fluctuates due to its less spherical shape compared to a 14-panel
seamless and stitch-less ball.
A small inflate pressure variation close to its recommended pressure has negligible effect on aerodynamic
properties.
Acknowledgements
The authors are highly grateful to the Australian Football Federation for providing the balls for this study and
their strong support.
References
[1] Alam F, Subic A, Watkins S, Naser J, Rasul M G. An experimental and computational study of aerodynamic properties of rugby balls.
WSEAS Transactions on Fluid Mechanics 2008;3:279-286.
[2] Alam F, Subic A, Watkins S, Smits A J. Aerodynamics of an Australian rules foot ball and rugby ball. Computational Science and
Engineering (edited by M. Peters), Springer, Germany; 2009.
[3] Alam F, Zimmer G, Watkins S. Mean and time-varying flow measurements on the surface of a family of idealized road vehicles. Journal
of Experimental Thermal and Fluid Sciences 2003;27:639-654.
[4] Mehta R D, Alam F, Subic A. Aerodynamics of tennis balls- a review. Sports Technology 2008;1:1-10.
[5] Smits A J, Ogg S. Golf ball aerodynamics. The Engineering of Sport 5, 2004;1:3-12.
[6] Asai T, Carré M J, Akatsuka T, Haake S J. The Curve Kick of a Football. Sports Engineering 2002;5:183–192.
[7] Asai T, Akatsuka T, Haake S J. The physics of football. Physics World 1998;11:25–27
[8] Barber S, Chin S B, Carre M J. Sports aerodynamics: A numerical study of the erratic motion of soccer balls. Computer & Fluids 2009;
38:1091-1100.
[9] Asai T, Seo K, Kobayashi O, Sakashita R. Fundamental aerodynamics of the soccer ball. Sports Engineering 2007:10:101-110.
2448 F. Alam et al. / Procedia Engineering 2 (2010) 2443–2448
... The ball traveling trajectory of a successful direct free kick likes a banana so that the ball flies over the player wall to the goal. A model predicting a banana free-kick must include both the effects of gravity, aerodynamic characteristics (such as the air resistance), and Magnus [3][4][5][6][7][8]. The effects of the initial velocity, kick force, and kick position on the ball traveling trajectory were investigated [9][10][11][12]. ...
... The dynamics model of the soccer ball is established under the four assumptions[13]: (1) the air resistance applied on the ball to be homogenous, (2) ignoring the ball flying spin,(3) ignoring the Magnus effect on the ball, (4) ignoring the gyroscopic moment of the ball, and (5) neglecting acceleration of Coriolis. Then, Equations of Motion 3.1-3.3 ...
Chapter
In this Chapter, the dynamics modeling, and virtual simulation of soccer ball free kick are investigated. A three body dynamics model is created based on the Newton second law and Hooke’s law. The model includes gravity, aerodynamic drag, Magnus force, and contact force. The motion of soccer ball is established as the time-dependent ordinary differential equations (ODEs). The virtual prototype technology is introduced and applied for the dynamic simulation of soccer ball free kicks. Some typical results are obtained.
... For example, the effects of the projectile angle on a particle traveling trajectory, velocity, and acceleration are studied using the equation of motion with consideration of air resistance [10]. Additionally, the effects of initial velocity, kick force, and kick position on the projectile trajectory of a soccer ball are explored [11,12,8]. ...
... The integration of three sub-models established in Sections 2.1-2.3 builds a dynamics simulation system for capturing kinematics and dynamics performance of a soccer ball projectile motion. Using a tool of dynamics software [11], it can The solution for these models supports the sports analysis of soccer ball. The dynamics models can help players to find the kicking force, kicking position and kicking angle to achieve an expected trajectory. ...
Chapter
In this Chapter, a spatial two-body dynamics model of soccer field-ball has been developed. The model involves gravity, aerodynamic drag, and contact force. The trajectory of the ball flight and the force applied to the ball are visualized through virtual kinematics and dynamics simulation. All works form a complete procedure to illustrate the application of dynamic simulation on soccer ball projectile motion analysis.
... The ball traveling trajectory of a successful direct free kick likes a banana so that the ball flies over the player wall to the goal. A model predicting a banana free-kick must include both the effects of gravity, aerodynamic characteristics (such as the air resistance), and Magnus [3][4][5][6][7][8]. The effects of the initial velocity, kick force, and kick position on the ball traveling trajectory were investigated [9][10][11][12]. ...
... The dynamics model of the soccer ball is established under the four assumptions[13]: (1) the air resistance applied on the ball to be homogenous, (2) ignoring the ball flying spin,(3) ignoring the Magnus effect on the ball, (4) ignoring the gyroscopic moment of the ball, and (5) neglecting acceleration of Coriolis. Then, Equations of Motion 3.1-3.3 ...
Chapter
Full-text available
The intelligent analysis of the motion of a soccer ball requires accurately predicting its trajectory and finding the accurate design parameters to achieve a goal. The optimization design can efficiently improve the design parameters within the required range to exact evaluation of flight trajectory. In this Chapter, an optimization method is proposed to plan, evaluate, and optimize the traveling trajectory of a soccer ball. The parameterization technology has been integrated into the optimization design for automatically capturing the expected target, which is expressed as a function of all design parameters. A case study goes through multi-body dynamics modeling, dynamic simulation, and optimal objective modeling.
... For example, the effects of the projectile angle on a particle traveling trajectory, velocity, and acceleration are studied using the equation of motion with consideration of air resistance [10]. Additionally, the effects of initial velocity, kick force, and kick position on the projectile trajectory of a soccer ball are explored [11,12,8]. ...
... The integration of three sub-models established in Sections 2.1-2.3 builds a dynamics simulation system for capturing kinematics and dynamics performance of a soccer ball projectile motion. Using a tool of dynamics software [11], it can The solution for these models supports the sports analysis of soccer ball. The dynamics models can help players to find the kicking force, kicking position and kicking angle to achieve an expected trajectory. ...
Chapter
这The kinematics and dynamics modeling are an important factor to realize intelligent analysis of a soccer ball motion in a virtual environment. The problem of how to accurately create a three-dimensional (3D) soccer model is addressed in this Chapter. The main steps are introduced to build a basic dynamics model of soccer ball. A numerical validation method is suggested to validate the dynamics models of soccer ball. The virtual prototype technology is applied for the solution of model validation problems in the quest for using a virtual simulation instead of a physical simulation. A case study illustrates the dynamics model is highly reliable in predicting the flight trajectory of a soccer ball.
... Their results show that at 30 m s −1 and 0 • seam angle, lift force, which acts perpendicular to the swing and drag forces, increases from less than 0.1 N at 0 Hz rotation rate, to between 0.3 and 0.45 N at 8.3-16.7 Hz, with these forces only 'weakly dependent on the speed of rotation'. These results are explained by the Magnus effect which causes balls in tennis, baseball, football and golf to curve in the air, as described in Mehta & Pallis (2001), Alam et al. (2010aAlam et al. ( , 2012, Bearman & Harvey (1976) and Mehta (1985). Mehta (2014) discusses how slow bowlers, as well as unusual side-arm bowlers, are able to use the Magnus effect to make cricket balls deviate laterally, but these phenomena are not the focus of this review. ...
Article
Full-text available
This paper reviews 70 years of cricket ball swing literature and reassesses the results in light of new measurements. A comparison of ball tracking data with published experimental results shows that current understanding does not explain the behaviour observed in professional cricket matches. Descriptions of cricket ball boundary layer aerodynamics are updated with new results which show that the seam acts more like a series of vortex generators than a trip and that there is a laminar separation bubble on the seam side of the ball. Previous results for reverse swing are consolidated and compared with new data, showing that different magnitudes of swing occur depending on the condition of both sides of the ball. Variation in pressure and temperature should be included when the Reynolds number, a non-dimensional ball speed, is calculated, while humidity can be neglected. Studies on the effect of wind speed and direction are summarised and results considering the effect of free-stream turbulence are compared with new measurements. Throughout the paper, recommendations for future work are suggested. These include quantitative study of the effect of surface defects and roughness, an assessment of whether atmospheric turbulence can affect swing and investigation into the effect of backspin on swing.
... It revealed that the Reynolds number and the total seam length had a strong correlation. Alam et al. [13][14][15] experimentally evaluated with aerodynamic forces and moments for several FIFA-approved soccer balls under different wind conditions. Another experimentation of similar nature was undertaken by Alam et al. [16]. ...
Article
Full-text available
The football game is the most popular, played, and loved sport around the world. The advent of technological breakthroughs and the continuous increase in consumer demand have led to a revolution in football’s design and manufacturing process. In the past, studies in soccer ball aerodynamics mainly were limited to the investigation of lift and drag forces inside a wind tunnel apparatus. A few researchers have analyzed the flow around the different soccer balls using computational fluid dynamics simulations with the Reynolds-Averaged-Navier–Stokes equations model. This study primarily intends to simulate a modern soccer ball (Adidas Telstar 18) using the Large Eddy Simulations technique. The whole research is divided into two phases. In the first phase, the flow around a smooth sphere is simulated numerically to validate the meshing strategy, boundary conditions, and solution methodology. The same modeling approach is used in the later stage to simulate the flow around a soccer ball. The effect of panels and seam on the boundary layer flow separation and overall turbulent flow structure around the soccer ball are visualized. The results indicate that the large-eddy simulations help predict the flow intricacies by resolving small eddies near the panels.
... We simulated different trajectories under two different conditions: gravity only and gravity + air drag. Air viscosity around the ball (ρ) was set to 1.225 kg/m 3 (value at sea level), and C d was set to 0.346 or 0.4 for baseball and soccer balls, respectively (Alam et al., 2010;Kagan and Nathan, 2014). We introduced one initial vertical speed V y 0 = 9.807 m/s (2 s of flight time under gravity only conditions) for each trajectory and corresponding approaching speeds (V y 0 = 7.5, 15, 25 m/s) different balls launched at the origin. ...
Article
Full-text available
Catching a ball in a parabolic flight is a complex task in which the time and area of interception are strongly coupled, making interception possible for a short period. Although this makes the estimation of time-to-contact (TTC) from visual information in parabolic trajectories very useful, previous attempts to explain our precision in interceptive tasks circumvent the need to estimate TTC to guide our action. Obtaining TTC from optical variables alone in parabolic trajectories would imply very complex transformations from 2D retinal images to a 3D layout. We propose based on previous work and show by using simulations that exploiting prior distributions of gravity and known physical size makes these transformations much simpler, enabling predictive capacities from minimal early visual information. Optical information is inherently ambiguous, and therefore, it is necessary to explain how these prior distributions generate predictions. Here is where the role of prior information comes into play: it could help to interpret and calibrate visual information to yield meaningful predictions of the remaining TTC. The objective of this work is: (1) to describe the primary sources of information available to the observer in parabolic trajectories; (2) unveil how prior information can be used to disambiguate the sources of visual information within a Bayesian encoding-decoding framework; (3) show that such predictions might be robust against complex dynamic environments; and (4) indicate future lines of research to scrutinize the role of prior knowledge calibrating visual information and prediction for action control.
... Based on the sphere diameters and that the measurement was taken at 20˚C, the Re range is between 6 x 10 3 and about 2.9 x 10 5 and the C d value is reported to be around 0.45-0.50 [39,53,54]. Using the wind tunnel setup described above, steady sphere drag was measured through the wind velocity range of 1 to 22 m s -1 . ...
Article
Full-text available
Vascular epiphytes represent almost 10% of all terrestrial plant diversity. Being structurally dependent on trees, epiphytes live at the interface of vegetation and atmosphere, making them susceptible to atmospheric changes. Despite the extensive research on vascular epiphytes, little is known about wind disturbance on these plants. Therefore, this study investigated the wind-epiphyte mechanical interactions by quantifying the drag forces on epiphytic bromeliads when subjected to increasing wind speeds (5–22 m s ⁻¹ ) in a wind tunnel. Drag coefficients ( C d ) and Vogel exponents ( B ) were calculated to quantify the streamlining ability of different bromeliad species. Bromeliads’ reconfiguration occurred first via bending and aligning leaves in the flow direction. Then leaves clustered and reduced the overall plant frontal area. This reconfiguration caused drag forces to increase at a slower rate as wind velocity increased. In the extreme case, drag force was reduced by 50% in a large Guzmania monostachia individual at a wind velocity of 22 m s ⁻¹ , compared to a stiff model. This species had one of the smallest C d (0.58) at the highest wind velocity, and the largest negative mean B (-0.98), representing the largest reconfiguration capacity amongst the tested bromeliads. The streamlining ability of bromeliads was mainly restricted by the rigidity of the lower part of the plant where the leaves are already densely clustered. Wind speeds used in this study were generally low as compared to storm force winds. At these low wind speeds, reconfiguration was an effective mechanism for drag reduction in bromeliads. This mechanism is likely to lose its effectiveness at higher wind speeds when continuous vigorous fluttering results in leaf damage and aspects such as root-attachment strength and substrate stability become more relevant. This study is a first step towards an understanding of the mechanical bottleneck in the epiphyte-tree-system under wind stress.
Article
Full-text available
The paper describes the critical Reynolds number of soccer balls as a function of the total groove volume on the ball surface. In this study, the aerodynamic drag and separation point of 6 soccer balls with different panel orientations were measured in a wind tunnel. Moreover, the total volume of grooves on the ball’s surface was measured by One shot 3D measuring machine. It was found that the drag coefficient in the supercritical region and the critical Reynolds number depend on the design of the ball surface (shape, number of panels, surface roughness). A negative correlation exists between the total groove volume on the surface and the critical Reynolds number. When the total groove volume is large, the drag in the supercritical region increases. The position of the separation point does not change in the subcritical and supercritical regions even if the orientation of the panel is changed. However, when the groove width is wider, the panel orientation dependence becomes significant in the supercritical region. In the transition region, the position of the separation point differs depending on the panel's orientation.
Article
Wind-tunnel experiments were performed on Al Rihla, Telstar 18, Brazuca, and Jabulani, official match soccer balls used in the World Cup in 2022, 2018, 2014, and 2010, respectively. Aerodynamic coefficients extracted from wind-tunnel results were used to create computer simulations of trajectories of kicked soccer balls. The newest World Cup ball, Al Rihla, could be kicked for longer range compared to its recent predecessors. But there is also the possibility that knuckling effects at large speeds could be seen in the flights of the new ball kicked with little to no spin. Measurements were made of each soccer ball’s total seam length and seam width and depth. Two different measurement techniques were employed to assist with accuracy. Wider and deeper seams led to smaller critical speeds.
Article
Full-text available
The aerodynamic properties of rugby balls as a function of wind speeds and yaw angles were measured using experimental and computational methods. The average drag and side forces coefficients for all speeds and yaw angles were computed and compared. The flow pattern around the rugby ball was visualized using wool tuft and smoke. No significant variation in Reynolds numbers was found in CFD results, however, some variations were noted in experimental findings. The average drag coefficient of a rugby ball at zero yaw was 0.18 and 0.14 in experimental and computational studies and this rose to about 0.60 and 0.50 when yawed at 90 degrees respectively.
Article
Full-text available
When the boundary layer of a sports ball undergoes the transition from laminar to turbulent flow, a drag crisis occurs whereby the drag coefficient (C d) rapidly decreases. However, the aerodynamic properties and boundary-layer dynamics of a soccer ball are not yet well understood. In this study we showed that the critical Reynolds number (Re crit) of soccer balls ranged from 2.2 × 105 to 3.0 × 105. Wind-tunnel testing, along with visualisation of the dynamics of the boundary layer and the trailing vortex of a ball in flight, demonstrated that both non-spinning and spinning (curved) balls had lowC d values in the super-critical region. In addition, theRe crit values of the soccer balls were lower than those of smooth spheres, ranging from ∼ 3.5 × 105 to 4.0 × 105, due to the effects of their panels. This indicated that the aerodynamic properties of a soccer ball were intermediate between those of a smooth ball and a golf ball. In a flow visualisation experiment, the separation point retreated and theC d decreased in a super-critical regime compared with those in a sub-critical regime, suggesting a phenomenon similar to that observed in other sports balls. With some non-spinning and spinning soccer balls, the wake varied over time. In general, the high-frequency component of an eddy dissipated, while the low-frequency component increased as the downstream vortex increased. The causes of the large-scale fluctuations in the vortex observed in the present study were unclear; however, it is possible that a ‘knuckle-ball effect’ of the non-rotating ball played a role in this phenomenon.
Article
Full-text available
The application of the commercial CFD code, FLUENT, to sports ball aerodynamics was assessed and a validated 3D analysis technique was established for balls that have been scanned with a 3D laser scanner or drawn in CAD. The technique was used to examine the effects of surface geometry on the aerodynamic behaviour of soccer balls by comparing the flow around different balls and predicting the aerodynamic force coefficients. The validation process included performing CFD studies on 3D smooth spheres and various soccer balls, and comparing the results to wind tunnel tests and flow visualisation. The CFD technique used a surface wrapping meshing method and the Reynolds Averaged Navier–Stokes approach with the realizable k-ε turbulence model, which was found to be able to predict the drag, lift and side force coefficients (CD, CL and CS) reliably, to compare the wake behaviour, and to give good pressure distributions near the stagnation point. The main limitations of the technique with the available computational resources were its inability to accurately predict boundary layer transition or growth, but despite this, several conclusions could be drawn regarding soccer ball aerodynamics. CD was not significantly different between balls. CL and CS were found to be significantly affected by the orientation of the ball relative to its direction of travel, meaning that balls kicked with low amounts of spin could experience quasi-steady lift and side forces and move erratically from side-to-side or up and down through the air. For different balls, CD, CL and CS were predicted and their variation with orientation entered into a modified trajectory simulation program. The erratic nature of this type of kick was found to vary with details of the surface geometry including seam size, panel symmetry, number, frequency and pattern, as well as the velocity and spin applied to the ball by the player. Exploitation of this phenomenon by players and ball designers could have a significant impact on the game.
Article
Bill Shankly, the former manager of Liverpool football club, once said: "Football is not about life or death. It is more important than that." This month at the World Cup in France, millions of football fans will get that same feeling for a few short weeks. Then the event will be over, and all that will remain will be a few repeats on television and the endless speculation about what might have happened. It is this aspect of football that its fans love, and others hate. What if that penalty had gone in? What if the player hadn't been sent off? What if that free kick hadn't bent around the wall and gone in for a goal?
Article
The aerodynamics of a tennis ball are reviewed here with reference to several wind tunnel measurement efforts. Measurements for a wide variety of tennis balls, including the ‘oversized’ balls, are presented. Flow visualization results have shown that the separation location on a non‐spinning tennis ball occurred relatively early, near the apex, and appeared very similar to a laminar separation in the subcritical Reynolds number regime. The flow regime (boundary layer separation location) appears to be independent of Reynolds number in the range, 167,000 < Re < 284,000. Asymmetric boundary layer separation and a deflected wake flow, depicting the Magnus effect, have been observed for the spinning ball. Aerodynamic force (drag and lift) measurements for nonspinning and spinning balls are reviewed for a wide range of Reynolds numbers and spin rates. Relatively high drag coefficients (CD ≅ 0.6 to 0.7), have been measured for new nonspinning tennis balls. The observed (unexpected) behavior of the tennis ball drag coefficient is explained in terms of a flow model that includes the drag contribution of the fuzz’ elements.
Article
The purpose of this study is to assess the fundamental characteristics which cause a football to spin in a curve ball kick. The ball impact process was analysed initially with a high speed video camera running at 4500 frames per second to obtain the basic data for a computer simulation model. This simulation model showed suitable agreement although it slightly deteriorated during the latter half of impact. It was noted that rotation of the ball occurs, even if the kinetic coefficient of friction is nearly equal to 0 because of local deformation of the ball during impact around the foot allowing forces to be transmitted to the ball around its axis. The spin of the ball was found to increase with the offset distance between the foot and the axis of the ball and as the kinetic coefficient of friction was increased. The offset distance between the foot and the axis of the ball affects the spin more than the coefficient of friction. Varying the coefficient of friction from 0.0 to 1.0 produces an increase in spin of 13 rad s−1 at most. It was suggested that the most suitable offset distance, which makes the largest ball rotation was around 100 mm. A trade-off was found between the ball speed and spin, for different offset distances.
Chapter
The aerodynamic behavior of a Rugby ball and an Australian Rules foot ball is complex and significantly differs from spherical sports balls due to their complex ellipsoidal shapes. Although prior aerodynamic studies have been conducted on soccer, tennis, cricket and golf balls, scant information about the Australian Rules and Rugby balls is available in the public domain. In order to understand the aerodynamic properties of Rugby and Australian Rules foot balls, experimental and computation studies have been undertaken for a range of speeds and yaw angles. The airflow around Rugby and Australian Rules foot balls was visualized and the average drag coefficients for both balls were determined and compared. Minor Reynolds number sensitivity at zero yaw angle was found in the experimental studies for both balls. However, significant Reynolds number variations were noted at yaw angles between 75∘ and 85∘. In contrast, no major Reynolds number dependency was found in the Computational Fluid Dynamics (CFD) results for both balls.
Article
Aerodynamically generated noise in the A-pillar region of a passenger vehicle can make a significant contribution to interior noise and adversely affect passenger comfort. Modern vehicles experience strong surface pressure fluctuations on the side windows due to flow separation around the A-pillar. The objective of this work is to investigate how these fluctuations vary with speeds and also to study the effects of local A-pillar and windshield geometry and yaw angles on the nature of the fluctuations. Pressure fluctuations in cavities (where strong feedback mechanisms can occur) were not studied in this work. Investigations were performed to study the effects of the A-pillar and windshield effective corner radii on the local flow and its potential for noise generation using a group of idealized road vehicle models. Surface mean and fluctuating pressures were measured on the side window in the A-pillar region at different Reynolds numbers and yaw angles. Tests were carried out in the RMIT University Industrial Wind Tunnel. Flow structure was documented using flow visualization. The studies show that Reynolds number sensitivities (based on free-stream flow velocities and body width) in the range of 6.4×105–1.5×106 were minimal. Surface mean and fluctuating pressures scaled well with velocity head and Strouhal number. The magnitudes of the fluctuating pressure coefficients depend largely on local A-pillar radii and can be reduced significantly with the increase of local corner radii.