ArticlePDF Available

Effect of Surface Groove Structure on the Aerodynamics of Soccer Balls

Authors:

Abstract and Figures

Soccer balls have undergone dramatic changes in their surface structure that can affect their aerodynamics. The properties of the soccer ball surface such as the panel shape, panel orientation, seam characteristics, and surface roughness have a significant impact on its aerodynamics and flight trajectory. In this study, we performed wind-tunnel tests to investigate how the introduction of grooves on the surface of a soccer ball affects the flight stability and aerodynamic forces on the ball. Our results show that for soccer balls without grooves, changing the panel orientation of the ball causes a significant change in the drag coefficient. Soccer balls with grooves exhibited a smaller change in air resistance (Cd) in the supercritical region (20 to 30 m/s; 3.0 × 10 5 ≤ Re ≤ 4.7 × 10 5), compared to the ungrooved ball where only the panel orientation was changed. Furthermore, at power-shot speeds (25 m/s), the grooved ball exhibited smaller variations in lift force and side force than the ungrooved ball. These results suggest that a long groove structure on the surface of the soccer ball has a significant impact on the air flow around the ball in the supercritical region, and has the effect of keeping the air flow separation line constant.
Content may be subject to copyright.
applied
sciences
Article
Eect of Surface Groove Structure on the
Aerodynamics of Soccer Balls
Sungchan Hong 1, 2, * and Takeshi Asai 1,2
1Faculty of Health and Sport Sciences, University of Tsukuba, Tsukuba 305-8574, Japan;
asai.takeshi.gf@u.tsukuba.ac.jp
2
Advanced Research Initiative for Human High Performance, University of Tsukuba, Tsukuba 305-8574, Japan
*Correspondence: hong.sungchan.fu@u.tsukuba.ac.jp; Tel.: +81-29-853-2650
Received: 17 July 2020; Accepted: 24 August 2020; Published: 25 August 2020


Abstract:
Soccer balls have undergone dramatic changes in their surface structure that can aect their
aerodynamics. The properties of the soccer ball surface such as the panel shape, panel orientation,
seam characteristics, and surface roughness have a significant impact on its aerodynamics and flight
trajectory. In this study, we performed wind-tunnel tests to investigate how the introduction of
grooves on the surface of a soccer ball aects the flight stability and aerodynamic forces on the ball.
Our results show that for soccer balls without grooves, changing the panel orientation of the ball
causes a significant change in the drag coecient. Soccer balls with grooves exhibited a smaller change
in air resistance (Cd) in the supercritical region (20 to 30 m/s; 3.0
×
10
5
Re
4.7
×
10
5
), compared
to the ungrooved ball where only the panel orientation was changed. Furthermore, at power-shot
speeds (25 m/s), the grooved ball exhibited smaller variations in lift force and side force than the
ungrooved ball. These results suggest that a long groove structure on the surface of the soccer ball
has a significant impact on the air flow around the ball in the supercritical region, and has the eect of
keeping the air flow separation line constant.
Keywords: aerodynamics; groove structure; new design; seam structure; surface shape
1. Introduction
In recent years, a variety of patterns have been used to make the surface panels for soccer balls.
The panels have undergone dramatic changes in structure, with new types of soccer balls appearing
at every FIFA World Cup since the 2006 event in Germany through to the 2018 event in Russia.
Recently, in addition to the ocial World Cup balls, soccer balls with unique panel structures have
been manufactured by dierent sports manufacturers and have been adopted as the ocial balls in
dierent soccer leagues around the world. Changes to the number of panels that make up a soccer ball
have been reported to significantly aect the flight trajectory of the ball and subsequently influence
the performance of players [
1
4
]. Similarly, it has been reported that the seam characteristics (length,
depth, and width) of the panel surface also have a significant impact on the aerodynamics and flight
trajectory [
5
]. Furthermore, the direction and number of seams can change the air flow around the
soccer ball and move the position of the air flow separation line [
5
,
6
]. Researchers have also analyzed
the aerodynamics and flight trajectories of balls from various other sports using a variety of techniques
to investigate the eect of factors such as the roughness (bumpy dimples) of the ball’s surface [716].
However, the eect of design changes on aerodynamic forces, due to changes in the structure of the
surface of the ball, requires further investigation. In this study, we investigated how the aerodynamics
of an existing soccer ball (TUJI-FA, Nassau, four-panel) can be altered by directly introducing a groove
that is shallower than the seams on the surface of the ball. In this study, conducted as basic research on
the impact of grooves, the aerodynamics were examined using a soccer ball with large panels, so that
Appl. Sci. 2020,10, 5877; doi:10.3390/app10175877 www.mdpi.com/journal/applsci
Appl. Sci. 2020,10, 5877 2 of 8
new grooves could be easily introduced on the panel surface. We examined whether it is possible to
change the aerodynamics of the ball in a less intrusive manner by introducing a groove structure into
a ball with a fixed number of panels. We produced both new and old type (T1 and T2, respectively)
soccer balls with grooves evenly inserted between the panels on the ball surface, with no change in the
number of panels. We then compared an existing old-type (TUJI-FA, four-panel, T1) ball to a new-type
(four-panel, T2) ball with the added groove structure, focusing on the aerodynamics of the balls and
the variation in aerodynamic force due to dierences in the orientation of the panels. In this study,
the test focused on a fixed non-rotating soccer ball. It should be noted that measuring a rotating soccer
ball is also important and will be part of a future study; however, in this case, a non-rotating ball was
examined to obtain the initial basic and fundamental data.
2. Methods
Wind-Tunnel Experiment
For this study, we used a re-circulating wind tunnel (San Technologies Co., Ltd., Tochigi, Japan)
located at the University of Tsukuba (Figure 1). The maximum wind velocity of the tunnel was 55 m/s,
the outlet dimensions were 1.5 m
×
1.5 m, the wind velocity distribution was within
±
0.5%, and the
turbulence was less than or equal to 0.1%. The wind tunnel system was able to automatically measure
the wind velocity at 0.1 m/s intervals from the Pitot-static tube located above the measurement position
of the soccer ball. To ensure that the flow generated from the Pitot-static tube would have no direct eect
on the flow around the ball, the ball was positioned at the exact center of the nozzle cross section and
the distance between the nozzle and the ball was adjusted to zero for the measurements. For example,
when the wind speed was set to 25 m/s, the measured mean wind speed was 25.28 m/s, with a deviation
from the measurement position in the range
0.46% to 0.45% and a wind speed distribution within
±
0.5%. Similarly, the degree of turbulence downstream of the nozzle when the wind speed was 25 m/s
was 0.05 to 0.06, which was within approximately
±
0.1%, so the error from the ball position had little
eect on the wind speed. Furthermore, the measuring system can automatically measure the dynamic
pressure at 0.1 Pa intervals by means of the Pitot-static tube placed above the measuring portion of the
soccer ball. The length of the sting used in this study was 0.6 m, and its width was 0.02 m. All the
soccer balls used in this study had the same diameter of 0.22 m, weight of 0.429
±
0.001 kg, and internal
pressure of 0.9 kgf/cm
2
. To measure the force of the soccer balls, we attached a round-shaped plate
to the rear of the soccer ball to fix it and connected it to the sting. In this arrangement, there was no
imbalance caused by changes such as distortions from a round sphere.
Appl. Sci. 2020, 10, x FOR PEER REVIEW 3 of 8
Figure 1. Wind-tunnel setup.
An old-type soccer ball (TUJI-FA, Nassau, Figure 2a) and a new-type soccer ball (Nassau, Figure
2c) were mounted on the wind tunnel for the experiments. The new soccer ball has the same four-
panel structure as the old ball. However, a groove in the pattern was introduced into the new-type
soccer ball (Figure 2b).
Figure 2. Soccer balls used in this experiment: (a) old-type soccer ball (T1), (b) shape of the groove
introduced into the ball surface (at four locations), and (c) new-type soccer ball (T2) with grooves.
The seams on the surface of the two soccer balls used in this experiment were 10 mm wide and
3.8 mm deep. The groove introduced into the new soccer ball (T2) was 450 mm long (450 × 4 = 1800
mm total length at all four locations), 5 mm wide, and 2 mm deep (Figure 3). The grooves were carved
on the ball surface using a press machine. The difference between a seam and a groove is that a seam
is generated between two panels that are tied together, and, depending on the soccer ball type, they
are sewn with thread or bond. A groove is a pattern introduced directly to the surface of a panel; it
has a unique shape with a slightly different depth and width from that of a seam.
Figure 3. Seam and groove structure: (a) the width and depth of the seam and (b) the groove structure.
The aerodynamic forces on the ball were measured at two different panel orientations, i.e., A
and B (Figure 4). Panel orientation (A) positions the air valve front and center, while panel orientation
Figure 1. Wind-tunnel setup.
Appl. Sci. 2020,10, 5877 3 of 8
An old-type soccer ball (TUJI-FA, Nassau, Figure 2a) and a new-type soccer ball (Nassau, Figure 2c)
were mounted on the wind tunnel for the experiments. The new soccer ball has the same four-panel
structure as the old ball. However, a groove in the pattern was introduced into the new-type soccer
ball (Figure 2b).
Appl. Sci. 2020, 10, x FOR PEER REVIEW 3 of 8
Figure 1. Wind-tunnel setup.
An old-type soccer ball (TUJI-FA, Nassau, Figure 2a) and a new-type soccer ball (Nassau, Figure
2c) were mounted on the wind tunnel for the experiments. The new soccer ball has the same four-
panel structure as the old ball. However, a groove in the pattern was introduced into the new-type
soccer ball (Figure 2b).
Figure 2. Soccer balls used in this experiment: (a) old-type soccer ball (T1), (b) shape of the groove
introduced into the ball surface (at four locations), and (c) new-type soccer ball (T2) with grooves.
The seams on the surface of the two soccer balls used in this experiment were 10 mm wide and
3.8 mm deep. The groove introduced into the new soccer ball (T2) was 450 mm long (450 × 4 = 1800
mm total length at all four locations), 5 mm wide, and 2 mm deep (Figure 3). The grooves were carved
on the ball surface using a press machine. The difference between a seam and a groove is that a seam
is generated between two panels that are tied together, and, depending on the soccer ball type, they
are sewn with thread or bond. A groove is a pattern introduced directly to the surface of a panel; it
has a unique shape with a slightly different depth and width from that of a seam.
Figure 3. Seam and groove structure: (a) the width and depth of the seam and (b) the groove structure.
The aerodynamic forces on the ball were measured at two different panel orientations, i.e., A
and B (Figure 4). Panel orientation (A) positions the air valve front and center, while panel orientation
Figure 2.
Soccer balls used in this experiment: (
a
) old-type soccer ball (T1), (
b
) shape of the groove
introduced into the ball surface (at four locations), and (c) new-type soccer ball (T2) with grooves.
The seams on the surface of the two soccer balls used in this experiment were 10 mm wide and
3.8 mm deep. The groove introduced into the new soccer ball (T2) was 450 mm long (
450 ×4=1800 mm
total length at all four locations), 5 mm wide, and 2 mm deep (Figure 3). The grooves were carved on
the ball surface using a press machine. The dierence between a seam and a groove is that a seam is
generated between two panels that are tied together, and, depending on the soccer ball type, they are
sewn with thread or bond. A groove is a pattern introduced directly to the surface of a panel; it has a
unique shape with a slightly dierent depth and width from that of a seam.
Appl. Sci. 2020, 10, x FOR PEER REVIEW 3 of 8
Figure 1. Wind-tunnel setup.
An old-type soccer ball (TUJI-FA, Nassau, Figure 2a) and a new-type soccer ball (Nassau, Figure
2c) were mounted on the wind tunnel for the experiments. The new soccer ball has the same four-
panel structure as the old ball. However, a groove in the pattern was introduced into the new-type
soccer ball (Figure 2b).
Figure 2. Soccer balls used in this experiment: (a) old-type soccer ball (T1), (b) shape of the groove
introduced into the ball surface (at four locations), and (c) new-type soccer ball (T2) with grooves.
The seams on the surface of the two soccer balls used in this experiment were 10 mm wide and
3.8 mm deep. The groove introduced into the new soccer ball (T2) was 450 mm long (450 × 4 = 1800
mm total length at all four locations), 5 mm wide, and 2 mm deep (Figure 3). The grooves were carved
on the ball surface using a press machine. The difference between a seam and a groove is that a seam
is generated between two panels that are tied together, and, depending on the soccer ball type, they
are sewn with thread or bond. A groove is a pattern introduced directly to the surface of a panel; it
has a unique shape with a slightly different depth and width from that of a seam.
Figure 3. Seam and groove structure: (a) the width and depth of the seam and (b) the groove structure.
The aerodynamic forces on the ball were measured at two different panel orientations, i.e., A
and B (Figure 4). Panel orientation (A) positions the air valve front and center, while panel orientation
Figure 3.
Seam and groove structure: (
a
) the width and depth of the seam and (
b
) the groove structure.
The aerodynamic forces on the ball were measured at two dierent panel orientations, i.e., A and
B (Figure 4). Panel orientation (A) positions the air valve front and center, while panel orientation (B)
(which is the opposite of orientation (A)) positions the point where the seams meet directly opposite
the air valve at the front and center. T1_A indicates the old Type 1 soccer ball positioned at orientation
A, and T1_B indicates the same ball positioned at orientation B. Similarly, T2_A indicates the new Type
2 soccer ball positioned at orientation A, and T2_B indicates the same ball positioned at orientation B
(Figure 4).
We measured the aerodynamic forces applied to each ball at 1 m/s intervals at a wind velocity (U)
ranging from 7 m/s (Re
1.0
×
10
5
) to 30 m/s (Re
4.7
×
10
5
), the speed interval most commonly used
by actual soccer players. The Reynolds number (Re) is a dimensionless number that is defined as the
ratio of inertial force to viscous force and can be expressed as
Re =UD/v(1)
where Uis the velocity (m/s), Dis the ball diameter (m), and vis the kinematic viscosity coefficient (m2/s).
Appl. Sci. 2020,10, 5877 4 of 8
Appl. Sci. 2020, 10, x FOR PEER REVIEW 4 of 8
(B) (which is the opposite of orientation (A)) positions the point where the seams meet directly
opposite the air valve at the front and center. T1_A indicates the old Type 1 soccer ball positioned at
orientation A, and T1_B indicates the same ball positioned at orientation B. Similarly, T2_A indicates
the new Type 2 soccer ball positioned at orientation A, and T2_B indicates the same ball positioned
at orientation B (Figure 4).
Figure 4. Soccer balls used in this experiment and their panel orientations.
We measured the aerodynamic forces applied to each ball at 1 m/s intervals at a wind velocity
(U) ranging from 7 m/s (Re 1.0 × 10
5
) to 30 m/s (Re 4.7 × 10
5
), the speed interval most commonly
used by actual soccer players. The Reynolds number (Re) is a dimensionless number that is defined
as the ratio of inertial force to viscous force and can be expressed as
Re = UD/v (1)
where U is the velocity (m/s), D is the ball diameter (m), and v is the kinematic viscosity coefficient
(m
2
/s).
Air forces acting on the ball were measured during a 10 s time interval by a sting-type six-
component force detector (model number LMC-61256 by Nissho Electric Works Co, Ltd.). The data
were recorded for 10 s using a PC equipped with a A/D converter board with a sampling rate of 1000
Hz. The aerodynamic forces measured in this experiment were converted into a drag coefficient (Cd),
lift coefficient (Cl), and side force coefficient (Cs), as shown in Equations (2)(4),
𝐶 2𝐷
𝜌𝑈
𝐴
(2)
𝐶 2𝐿
𝜌𝑈
𝐴
(3)
𝐶 2𝑆
𝜌𝑈
𝐴
(4)
where ρ is the air density, expressed as ρ = 1.2 kg/m
3
; U is the flow velocity; and A is the projected
area of the soccer ball, expressed as A = π × 0.11
2
= 0.038 m
2
.
Figure 4. Soccer balls used in this experiment and their panel orientations.
Air forces acting on the ball were measured during a 10 s time interval by a sting-type six-component
force detector (model number LMC-61256 by Nissho Electric Works Co, Ltd., Tokyo, Japan). The data
were recorded for 10 s using a PC equipped with a A/D converter board with a sampling rate of
1000 Hz. The aerodynamic forces measured in this experiment were converted into a drag coecient
(Cd), lift coecient (Cl), and side force coecient (Cs), as shown in Equations (2)–(4),
Cd=
2D
ρU2A(2)
Cl=
2L
ρU2A(3)
Cs=
2S
ρU2A(4)
where
ρ
is the air density, expressed as
ρ
=1.2 kg/m
3
;Uis the flow velocity; and Ais the projected area
of the soccer ball, expressed as A=π×0.112=0.038 m2.
3. Results and Discussion
3.1. Drag Coecient Variation by Ball Type
Figure 5shows the drag characteristic curves of each soccer ball obtained from the wind-tunnel
tests. We positioned T1 and T2 soccer balls with two panel orientations (A and B) (Figure 4) and
measured the drag coecient of each soccer ball at wind velocities ranging from 7 m/s to 30 m/s.
The drag characteristic curves represent the mean values over three trials for each soccer ball. For each
ball, the Pearson product-moment correlation coecient for the results of the three trials was
r=0.95
(
p<0.01
), suggesting that there was no significant dierence between trials. In the case of the four-panel
T1 with no grooves, it was observed that changing the panel orientation caused a significant change in
the drag exerted on the ball. However, in the case of the T2 ball with new grooves introduced on the
surface of the panel, changing the panel orientation caused slight change in the drag exerted on the ball.
These drag coecient results suggest that introducing a shallow groove into the surface of the ball is an
eective way to minimize the variation in aerodynamic force due to changes in the position of the ball.
Appl. Sci. 2020,10, 5877 5 of 8
Appl. Sci. 2020, 10, x FOR PEER REVIEW 5 of 8
3. Results and Discussion
3.1. Drag Coefficient Variation by Ball Type
Figure 5 shows the drag characteristic curves of each soccer ball obtained from the wind-tunnel
tests. We positioned T1 and T2 soccer balls with two panel orientations (A and B) (Figure 4) and
measured the drag coefficient of each soccer ball at wind velocities ranging from 7 m/s to 30 m/s. The
drag characteristic curves represent the mean values over three trials for each soccer ball. For each
ball, the Pearson product-moment correlation coefficient for the results of the three trials was r = 0.95
(p < 0.01), suggesting that there was no significant difference between trials. In the case of the four-
panel T1 with no grooves, it was observed that changing the panel orientation caused a significant
change in the drag exerted on the ball. However, in the case of the T2 ball with new grooves
introduced on the surface of the panel, changing the panel orientation caused slight change in the
drag exerted on the ball. These drag coefficient results suggest that introducing a shallow groove into
the surface of the ball is an effective way to minimize the variation in aerodynamic force due to
changes in the position of the ball.
The relationship between wind velocity and drag coefficient is illustrated. T1_A indicates the
drag coefficient on the Type 1 soccer ball positioned at orientation A, and T1_B indicates the same at
orientation B. Likewise for the drag coefficient on the Type 2 soccer ball (T2_A, T2_B)
We found no significant difference in the critical Reynolds number (Re
c
) of each soccer ball under
the different panel orientations. For the T1 soccer ball, the critical Reynolds number was ~2.4 × 10
5
(Cd 0.16) under orientation A (T1_A) and ~2.4 × 10
5
(Cd 0.16) under orientation B (T1_B). For the
T2 soccer ball, the critical Reynolds number was ~2.6 × 10
5
(Cd 0.15) under orientation A (T2_A) and
~2.5 × 10
5
(Cd 0.15) under orientation B (T2_B). It can be observed that the difference was
insignificant. However, in the case of T1, the drag coefficient values varied significantly depending
on the orientation of the ball in the supercritical region (above Re = 2.5 × 10
5
) where the wind velocity
is higher. The fact that the air resistance acting on the T1 soccer ball (T1_A & T1_B) varies greatly in
this high-speed segment (at wind velocities of 20 to 30 m/s) suggests that changes in the orientation
of the T1 soccer ball could cause significant changes in its flight. In contrast, the T2 ball (T2_A & T2_B)
with grooves on the surface of the ball exhibited little orientation-dependent variation in
aerodynamic force even at high velocity (Figure 5), which suggests that this ball will follow a more
stable flight trajectory with less orientation-dependent variation over the course of its flight.
Figure 5. Drag coefficient of the soccer balls.
Figure 5. Drag coecient of the soccer balls.
The relationship between wind velocity and drag coecient is illustrated. T1_A indicates the
drag coecient on the Type 1 soccer ball positioned at orientation A, and T1_B indicates the same at
orientation B. Likewise for the drag coecient on the Type 2 soccer ball (T2_A, T2_B)
We found no significant dierence in the critical Reynolds number (Re
c
) of each soccer ball under
the dierent panel orientations. For the T1 soccer ball, the critical Reynolds number was ~2.4
×
10
5
(Cd
0.16) under orientation A (T1_A) and ~2.4
×
10
5
(Cd
0.16) under orientation B (T1_B). For the
T2 soccer ball, the critical Reynolds number was ~2.6
×
10
5
(Cd
0.15) under orientation A (T2_A)
and ~2.5
×
10
5
(Cd
0.15) under orientation B (T2_B). It can be observed that the dierence was
insignificant. However, in the case of T1, the drag coecient values varied significantly depending on
the orientation of the ball in the supercritical region (above Re =2.5
×
10
5
) where the wind velocity is
higher. The fact that the air resistance acting on the T1 soccer ball (T1_A & T1_B) varies greatly in this
high-speed segment (at wind velocities of 20 to 30 m/s) suggests that changes in the orientation of the
T1 soccer ball could cause significant changes in its flight. In contrast, the T2 ball (T2_A & T2_B) with
grooves on the surface of the ball exhibited little orientation-dependent variation in aerodynamic force
even at high velocity (Figure 5), which suggests that this ball will follow a more stable flight trajectory
with less orientation-dependent variation over the course of its flight.
3.2. Changes in the in Lift and Side Forces over Time
Figure 6plots the changes in standard deviations for the side and lift forces on each soccer ball
for a 10 s duration at a constant wind velocity of 25 m/s. T1_A indicates the changes in the standard
deviations of the side and lift forces on the Type 1 soccer ball positioned at orientation A for 10 s,
and T1_B indicates the same at orientation B. Similar changes were observed for the side and lift forces
on the Type 2 soccer ball (T2_A, T2_B) (10 s, wind velocity of 25 m/s). When T1 and T2 are compared,
it is possible to see that the drag coecient of T1 is more direction-dependent, but no significant
dierence was observed when the lift and side forces were changed.
Non-stationary forces often exhibit irregular fluctuations; furthermore, the fluctuations in the lift
and side forces acting on the ball varied depending on the type of soccer ball as shown in Figure 6.
Specifically, we found that the irregular fluctuations for aerodynamic force diered more between
dierent orientations of the same ball than between dierent ball types (Table 1). The variations in the
lift and side forces were approximately 15% smaller (p<0.1) for the T2 ball with grooves compared to
the T1 ball, which suggests that introducing a shallow groove structure on the surface of a soccer ball
is an eective way to minimize the irregular vertical and horizontal force fluctuations and hence can
Appl. Sci. 2020,10, 5877 6 of 8
be expected to produce a more stable flight trajectory. Specifically, the results of Figure 6indicate that
the forces standard deviation (SD) vary more widely when the orientation of the same ball is changed
than when dierent ball types are used (Table 1).
Figure 6.
Scatter diagrams of changes in the lift and side forces acting on the balls in the wind tunnel test.
Table 1.
Mean and standard deviation of the lift and side forces on the soccer balls used in this
experiment (10 s at wind velocity of 25 m/s).
Type
T1_A T1_B T2_A T2_B
Side
Force
Lift
Force
Side
Force
Lift
Force
Side
Force
Lift
Force
Side
Force
Lift
Force
Mean 1.10 1.38 0.77 2.69 0.83 1.74 1.85 1.50
SD 1.23 1.05 0.70 0.80 0.92 0.88 0.74 0.70
Previous research has shown that adding a bumpy dimple pattern to the soccer ball’s surface
causes the air flow separation line to move backwards, thereby reducing the force (air resistance)
exerted on the ball [
17
]. It has also been reported that the aerodynamics of a soccer ball are significantly
aected by the characteristics of its seams, in particular the length, width, and depth of the seam [
1
].
With respect to the length of the seam in particular, it has been reported that the longer the seam,
the smaller the critical Reynolds number (Re
c
) [
1
,
2
]. However, in this study, the value of Re
c
for the
T1 soccer ball (with a seam length of 2300 mm) was ~2.4
×
10
5
(Cd
0.16) for T1_A and ~2.4
×
10
5
(Cd
0.16) for T1_B. For the T2 soccer ball with grooves (each 450 mm long) at four locations, with a
seam length of 2300 mm and a total groove length of 1800 mm, the value of Re
c
was ~2.6
×
10
5
(Cd
0.15) for T2_A and ~2.5
×
10
5
(Cd
0.15) for T2_B. The fact that the two balls exhibit little
dierence in the value of Re
c
, counter to the result from previous research [
1
,
2
] on the relationship
between the length of the seam and the critical Reynolds number (Re
c
). This is considered to be due to
the fact that the depth (2 mm) and width (5 mm) of the groove introduced in this study are shallower
and narrower than the depth (3.8 mm) and width (10 mm) of the seam. In addition, previous research
on golf balls has shown that the form of the dimples on the surface of the ball has a large impact on
the ball’s aerodynamics, and that the ball’s wind velocity range characteristics depend on the size
of the dimples [
18
]. We speculate that the deep seams and shallow grooves on the surface of the
soccer ball in this study act like dimples of dierent sizes on a golf ball. Furthermore, the fact that
the shallow, narrow groove reduces the air resistance in the high-velocity segment suggests that the
groove structure has a significant impact on the air flow around the ball in the supercritical region.
Appl. Sci. 2020,10, 5877 7 of 8
4. Conclusions
In our previous research on the aerodynamics of soccer balls, we examined how dierent panel
shapes and positions, the number and direction of panels [
2
,
12
], the characteristics of the seams, and the
shape of the surface impacted the ball’s aerodynamic [
3
5
]. Further, for the experiment, we used
existing soccer balls and made simple changes to the roughness of their surfaces by directly introducing
grooves into them. For this reason, we believe this experiment was based on the results of our previous
study, and the data obtained from it supports our research results concerning the surface structure of
soccer balls. In this study, we conducted wind-tunnel tests on soccer balls to investigate the impact of
the ball’s surface structure on its aerodynamics. We compared the aerodynamics of an actual Type
1 (T1) soccer ball with an actual Type 2 (T2) soccer ball that was modified by introducing a groove
into the surface of the ball. The wind-tunnel tests show that the ungrooved T1 soccer ball has a higher
level of dependence on the panel orientation, and that changes to the panel orientation can produce
significant changes in the drag exerted on the T1 ball. In contrast, the T2 ball with grooves on the
surface of the ball exhibited little orientation-dependent variation in drag, which suggests that this ball
delivers a more uniform and stable flight than the T1 ball. This is because the long and shallow grooves
on the surface of the T2 ball stabilize or regularize the air flow around the ball, thereby minimizing the
movement of the air flow separation line. In this study, the test was conducted on a non-rotating soccer
ball and the aerodynamic forces acting on a fixed soccer ball were measured. However, since rotating
balls are more frequent than non-rotating balls in actual matches, we intend to perform measurements
on a rotating soccer ball, using a wind tunnel, in future studies. Since this is a basic study that does
not consider the air flow, we are currently using a 3D CFD (Computational Fluid Dynamics) software
based on the lattice Boltzmann method [
19
] to analyze how these surface shapes aect the air flow on
the ball surface and the position of the detachment points. Additionally, we plan to use visualization
methods such as PIV (Particle Image Velocimetry) to conduct a more detailed investigation of how the
structure of the grooves on the surface of the soccer ball aects the structure of air flow around the ball
(generation, decay, and detachment) and the movement of the points of air flow separation.
Author Contributions:
Conceptualization, S.H. and T.A.; methodology, S.H.; software, S.H.; validation, S.H.;
formal analysis, S.H.; investigation, S.H.; resources, S.H. and T.A.; data curation, S.H. and T.A.; writing—original
draft preparation, S.H.; writing—review and editing, S.H.; funding acquisition, S.H. and T.A. All authors have
read and agreed to the published version of the manuscript.
Funding:
This research was funded by the Ministry of Education, Culture, Sports, Science and Technology of the
Japanese government, JSPS KAKENHI grant number 15K16442, 20K11413 and 20H04066.
Conflicts of Interest: The authors declare no conflict of interest.
References
1.
Naito, K.; Hong, S.; Koido, M.; Nakayama, M.; Sakamoto, K.; Asai, T. Eect of seam characteristics on critical
Reynolds number in footballs. Mech. Eng. J. 2018,5, 17-00369. [CrossRef]
2.
Hong, S.; Asai, T. Eect of panel shape of soccer ball on its flight characteristics. Sci. Rep.
2014
,4, 5068.
[CrossRef] [PubMed]
3.
Go, J.E.; Hong, S.; Asai, T. Aerodynamic and surface comparisons between Telstar 18 and Brazuca. Proc. Inst.
Mech. Eng. Part P J. Sports Eng. Technol. 2018,232, 342–348. [CrossRef]
4.
Go, J.E.; Asai, T.; Hong, S. A comparison of Jabulani and Brazuca non-spin aerodynamics. Proc. Inst. Mech.
Eng. Part P J. Sports Eng. Technol. 2014,228, 188–194. [CrossRef]
5.
Hong, S.; Asai, T.; Seo, K. Visualization of air flow around soccer ball using a particle image velocimetry.
Sci. Rep. 2015,5, 15108. [CrossRef] [PubMed]
6.
Go, J.E.; Hong, S.; Asai, T. Eect of a soccer ball’s seam geometry on its aerodynamics and trajectory.
Proc. Inst. Mech. Eng. Part P J. Sports Eng. Technol. 2019,234, 19–29. [CrossRef]
7.
Passmore, M.A.; Rogers, D.; Tuplin, S.; Harland, A.; Lucas, T.; Holmes, C. The aerodynamic performance of
a range of FIFA-approved footballs. Proc. Inst. Mech. Eng. Part P J. Sports Eng. Technol.
2011
,226, 61–70.
[CrossRef]
Appl. Sci. 2020,10, 5877 8 of 8
8.
Aoki, K.; Muto, K.; Okanaga, H. Aerodynamic characteristics and flow pattern of a golf ball with rotation.
Procedia Eng. 2010,2, 2431–2436. [CrossRef]
9.
Djamovski, V.; Rosette, P.; Chowdhury, H.; Alam, F.; Steiner, T. A comparative study of rugby ball
aerodynamics. Procedia Eng. 2012,34, 74–79. [CrossRef]
10.
Price, D.S.; Jones, R.; Harland, A.R. Computational modelling of manually stitched soccer balls. Proc. Inst.
Mech. Eng. Part L J. Mater. Des. Appl. 2006,220, 259–268. [CrossRef]
11.
Asai, T.; Ito, S.; Seo, K.; Hitotsubashi, A. Fundamental aerodynamics of a new volleyball. Sports Technol.
2010
,
3, 235–239. [CrossRef]
12.
Hong, S.; Weon, B.M.; Nakanishi, Y.; Kimachi, K.; Seo, K.; Asai, T. Aerodynamic eects of a panel orientation
in volleyball float serve. In Proceedings of the 2018 International Society of Biomechanics in Sports, Auckland,
New Zealand, 10–14 September 2018; pp. 1–4.
13.
Hong, S.; Asai, T.; Weon, B.M. Surface patterns for drag modification in volleyballs. Appl. Sci.
2019
,9, 4007.
[CrossRef]
14.
Asai, T.; Nakanishi, Y.; Akiyama, N.; Hong, S. Flow Visualization of Spinning and Nonspinning Soccer Balls
Using Computational Fluid Dynamics. Appl. Sci. 2020,10, 4543. [CrossRef]
15. Thomson, J.J. The dynamics of a golf ball. Nature 1910,85, 2151–2157.
16.
Mehta, R.D.; Bentley, K.; Proudlove, M.; Varty, P. Factors aecting cricket ball swing. Nature
1983
,303,
787–788. [CrossRef]
17.
Hong, S.; Asai, T. Aerodynamic eects of dimples on soccer ball surfaces. Heliyon
2017
,3, e00432. [CrossRef]
[PubMed]
18.
Naruo, T.; Mizota, T. The influence of golf ball dimples on aerodynamic characteristics. Procedia Eng.
2014
,
72, 780–785. [CrossRef]
19.
Chen, H. Volumetric formulation of the lattice Boltzmann method for fluid dynamics: Basic concept.
Phys. Rev. E 1998,58, 3955–3963. [CrossRef]
©
2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
... Many studies on the effects of the soccer balls' shapes on aerodynamics and flight trajectories have been conducted [2][3][4][5][6][7][8][9][10][11][12][13]. Studies on the effects of the panel orientation variations, owing to the number of panels reduction, on the aerodynamics and flight trajectories of the soccer balls have also been published [14][15][16][17][18]. In modern soccer balls, not only the number of panels but also the surface roughness (bumps and dimples) significantly change, substantially affecting the ball aerodynamics [1,[19][20][21][22]. ...
... In summary, the surface conditions (roughness, worn and torn leather materials) of the modern soccer balls change with usage, causing variations in the airflow around the balls, which affect the aerodynamics and flight trajectories of the balls. Based on the variations of the aerodynamic forces acting on the ball depending on the surface structure, we assumed that the surface conditions of the ball are essential to study their aerodynamics in addition to the number and shape of panels, ball orientation, and surface structure reported in previous studies [6,8,18]. ...
Article
Full-text available
The surface structure of soccer balls, such as the number and shapes of the ball panels, has recently changed, and research on the aerodynamics and flight trajectories of new soccer balls is actively proceeding. However, these studies are focused on new soccer balls, whereas the used soccer balls were never studied. In this study, the aerodynamic characteristics of soccer balls kicked 1000 times by a robot were investigated through wind tunnel tests. The results were compared with those obtained using new soccer balls. Regarding the aerodynamic characteristics of the soccer balls, it was found that the critical Reynold number, Recrit, changes with usage. This is related to the transition from laminar to turbulent flow of airflow around the ball. The comparison of the drag coefficients of the balls at Recrit showed that the drag coefficients of the new and used Telstar18 balls were 0.15 (Re = 2.5 × 105) and 0.14 (Re = 2.2 × 105), respectively; those of the new and used Merlin were 0.13 (Re = 2.8 × 105) and 0.13 (Re = 2.2 × 105), respectively; and finally, those of the new and used Derbystar were 0.14 (Re = 2.1×105) and 0.14 (Re = 2.1×105), respectively. The surface conditions of a soccer ball, such as the surface roughness and surface damages, are essential factors to determine the aerodynamics of the soccer balls.
Article
Material dan teknik pembuatan bola menentukan permukaan bola. Permukaan bola berpengaruh terhadap performansi bola dimana hal ini terkait dengan aspek aerodinamika yaitu perilaku bola di udara sekitar. Permukaan bola berpengaruh terhadap gerak bola di udara. Penelitian ini bertujuan untuk menganalisis struktur permukaan 5 merk bola sepak dan 5 merk bola voli. Lima merk bola sepak meliputi Adidas champion, Adidas telstar, Mikasa squadra, Nike academy, dan Padova. Lima merk bola voli meliputi Apollo, Nassau, Mikasa competition, Mikasa gold, Mikasa MG. Penelitian dilakukan dengan pengamatan permukaan bola menggunakan mikroskop digital USB dengan pembesaran 1600 kali. Pengolahan data dilakukan menggunakan aplikasi ImageJ versi 1.53k. Foto mikroskop di-crop untuk mendapatkan foto dengan ukuran 1 x 1 mm dan dikonversi dalam mode abu-abu 8 bit. Foto abu-abu ini selanjutnya digunakan untuk mendapatkan struktur permukaan dengan menggunakan metode plot surface. Hasil penelitian menunjukkan bahwa terdapat perbedaan struktur permukaan 5 merk bola sepak dan 5 merk bola voli.
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
Various studies have been conducted on the aerodynamic characteristics of nonspinning and spinning soccer balls. However, the vortex structures in the wake of the balls are almost unknown. One of the main computational fluid dynamics methods used for the analysis of vortex structures is the lattice Boltzmann method as it facilitates high-precision analysis. Studies to elucidate the dominant vortex structure are important because curled shots and passes involving spinning balls are frequently used in actual soccer games. In this study, we identify the large-scale dominant vortex structure of a soccer ball and investigate the stability of the structure using the lattice Boltzmann method, wind tunnel tests, and free-flight experiments. One of the dominant vortex structures in the wake of both nonspinning and spinning balls is a large-scale counter-rotating vortex pair. The side force acting on a spinning ball stabilizes when the fluctuation of the separation points of the ball is suppressed by the rotation of the ball. Thus, although a spinning soccer ball is deflected by the Magnus effect, its trajectory is regular and stable, suggesting that a spinning ball can be aimed accurately at the outset of its course.
Article
Full-text available
Surface patterns on objects are important in aerodynamics. We show how surface patterns on volleyballs modify their aerodynamic performances. Conventional volleyballs with six panels show different aerodynamic characteristics along transverse and diagonal directions. Interestingly, isotropic surface patterning with hexagons or dimples enables us to achieve isotropic aerodynamics. This result gives insight into surface-mediated flight controls of projectiles through resisting fluid media.
Article
Full-text available
Aerodynamic coefficients were determined for Telstar 18 and Brazuca, match balls for the 2018 and 2014 World Cups, respectively. Experimental determination of aerodynamic coefficients prompted the development of computationally determined soccer ball trajectories for most launch speeds experienced in actual play. Although Telstar 18’s horizontal range will be nearly 10% shorter than Brazuca’s horizontal range for high-speed kicks, both Telstar 18 and Brazuca have similar knuckling effects due to nearly equal critical speeds and high-speed drag coefficients that differ by less than 10%. Surface comparisons suggest why aerodynamic properties for the two World Cup balls are so similar.
Article
Full-text available
In recent years, the design of footballs, including the number and shape of the panels forming the surface of footballs, has undergone a significant change. However, panels of varied shapes and seams are combined in complex ways to form the surface shape of footballs, and almost nothing is known about the effect of these surface shapes on the football drag characteristics. The present study used a wind tunnel to study the relationship between the critical Reynolds number and the length, depth, and width of the panel joints (seams) of 10 most recent football types used in matches in recent years. The results show the tendency that the drag coefficient in the super-critical regime (30 m/s) of a football increases as the panel joint length increases. Moreover, the depth of the panel joint indicates the highest correlation with the critical Reynolds number (r = -0.71, p < 0.01) and is considered to be a strong and convenient indicator that expresses the roughness (large scale) of the football surface. This study reveals the drag characteristics of the latest footballs and enables some degrees of prediction of the critical Reynolds number for the latest football types and those that will be developed in the future.
Article
Full-text available
Aerodynamic forces act on a golf ball during flight. For that purpose in order to develop a golf ball with high performance, it is important to analyze aerodynamic characteristics of the golf ball. On the other hand, dimple pattern of a golf ball is complicate and the influence of dimples hasn’t been investigated in detail. Therefore the influence of golf ball dimples on aerodynamic characteristics was investigated using a wind tunnel and rotating device. Some golf balls whose dimples have different depths were measured. As a result, it was found that the shallower the dimple was, the larger the lift coefficient was. However, when the depth of the dimples was much shallower, the lift coefficient was extremely-little on the slow velocity, i.e. under 30m/s. When the golf ball trajectory which was launched with a driver was calculated under various initial conditions, the ball velocity became under 30 m/s over the vertex of the trajectory in many cases, including professional male golfers and female golfers. Therefore, if the lift coefficient of the velocity of 30m/s becomes smaller, distances will become shorter. Dimple patterns that had a high lift coefficient at all velocities was researched. As a result, it was found that a golf ball with a dimple pattern that has extremely small dimples between large shallow dimples has a high lift coefficient at all velocities, including under 30m/s. In order to investigate the cause of the results, a flow visualization experiment was conducted. Visualization of flow around a golf ball was conducted by generating smoke. Moreover pictures of flow were taken by high-speed video and analyzed by PIV (Particle Image Velocimetry). As a result, there was the difference in the streamline distribution between golf balls.
Article
Full-text available
Wind-tunnel experimental measurements of drag coefficients for non-spinning Jabulani and Brazuca balls are presented. The Brazuca ball's critical drag speed is lower than that of the Jabulani ball, and the Brazuca ball's super-critical drag coefficient is larger than that of the Jabulani ball. Compared to the Jabulani ball, the Brazuca ball suffers less instability due to knuckle-ball effects. Using drag data, numerically determined ball trajectories are created, and it is postulated that although power shots are too similar to note flight differences, goalkeepers are likely to note the differences between Jabulani and Brazuca ball trajectories for intermediate-speed ranges. This latter result may appear in the 2014 World Cup for goalkeepers used to the flight of the ball used in the 2010 World Cup.
Article
Full-text available
Soccer balls are typically constructed from 32 pentagonal and hexagonal panels. Recently, however, newer balls named Cafusa, Teamgeist 2, and Jabulani were respectively produced from 32, 14, and 8 panels with shapes and designs dramatically different from those of conventional balls. The newest type of ball, named Brazuca, was produced from six panels and will be used in the 2014 FIFA World Cup in Brazil. There have, however, been few studies on the aerodynamic properties of balls constructed from different numbers and shapes of panels. Hence, we used wind tunnel tests and a kick-robot to examine the relationship between the panel shape and orientation of modern soccer balls and their aerodynamic and flight characteristics. We observed a correlation between the wind tunnel test results and the actual ball trajectories, and also clarified how the panel characteristics affected the flight of the ball, which enabled prediction of the trajectory.
Article
Five different soccer balls, each possessing the traditional 32-panel surface design, were tested in a wind tunnel. Only seam depth and width varied between the balls. Wind-tunnel tests and an examination of correlation revealed that seam width with a linear fit R 2 = 0:90 and p \ 10 À4 was a stronger indicator of a ball's critical speed than seam depth. Wind-tunnel data were used for computational modeling of many soccer-ball trajectories. It was determined that variations in seam geometry resulted in fluctuations up to 4 m in the horizontal range of hard-hit, no-spin kicks that travel approximately 68 m. Those seam geometry variations also contributed to lateral deflections up to 4 m for the aforementioned hard-hit, no-spin kicks.
Article
Recently, the shape and design of the panel on the official ball used in the FIFA World Cup was considerably different from that of a conventional soccer ball (having 32 pentagonal and hexagonal panels). Depending on the number of different panels and their orientation, the aerodynamic force experienced by a ball is believed to change, which in turn changes the ball trajectory. However, not much is known about the impact of the surface forms of a ball on its aerodynamics. Therefore, in the present study, 10 different types of soccer balls were produced and their aerodynamic properties were studied by wind tunnel experiments. The results confirmed that the aerodynamic force acting on the ball varied considerably depending on the existence of dimples on the ball surface. In addition, the 4 types of soccer balls, which had different kinds of roughness, revealed that even balls having the same number and shapes of panels experienced greatly varying aerodynamic forces depending on the surface form of the balls.
Article
A traditional soccer ball is constructed using 32 pentagonal and hexagonal panels. In recent years, however, the likes of the Teamgeist and Jabulani balls, constructed from 14 and 8 panels, respectively, have entered the field, marking a significant departure from conventionality in terms of shape and design. Moreover, the recently introduced Brazuca ball features a new 6-panel design and has already been adopted by many soccer leagues. However, the shapes of the constituent panels of these balls differ substantially from those of conventional balls. Therefore, this study set out to investigate the flight and aerodynamic characteristics of different orientations of the soccer ball, which is constructed from panels of different shapes. A wind tunnel test showed substantial differences in the aerodynamic forces acting on the ball, depending on its orientation. Substantial differences were also observed in the aerodynamic forces acting on the ball in different directions, corresponding to its orientation and rotation. Moreover, two-dimensional particle image velocimetry (2D-PIV) measurements showed that the boundary separation varies depending on the orientation of the ball. Based on these results, we can conclude that the shape of the panels of a soccer ball substantially affects its flight trajectory.