Sunteți pe pagina 1din 10

Proceedings of the Institution of Mechanical

Engineers, Part P: Journal of Sports


Engineering and Technology
http://pip.sagepub.com/

Aerodynamics of a golf ball with grooves


Jooha Kim and Haecheon Choi
Proceedings of the Institution of Mechanical Engineers, Part P: Journal of Sports Engineering and Technology published
online 29 July 2014
DOI: 10.1177/1754337114543860

The online version of this article can be found at:


http://pip.sagepub.com/content/early/2014/07/21/1754337114543860
A more recent version of this article was published on - Nov 18, 2014

Published by:

http://www.sagepublications.com

On behalf of:

Institution of Mechanical Engineers

Additional services and information for Proceedings of the Institution of Mechanical Engineers, Part P: Journal of Sports Engineering and
Technology can be found at:

Email Alerts: http://pip.sagepub.com/cgi/alerts

Subscriptions: http://pip.sagepub.com/subscriptions

Reprints: http://www.sagepub.com/journalsReprints.nav

Permissions: http://www.sagepub.com/journalsPermissions.nav

Citations: http://pip.sagepub.com/content/early/2014/07/21/1754337114543860.refs.html

Version of Record - Nov 18, 2014

>> OnlineFirst Version of Record - Jul 29, 2014

What is This?

Downloaded from pip.sagepub.com at Monash University on December 6, 2014


Original Article

Proc IMechE Part P:


J Sports Engineering and Technology
19
Aerodynamics of a golf ball with IMechE 2014
Reprints and permissions:
grooves sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/1754337114543860
pip.sagepub.com

Jooha Kim and Haecheon Choi

Abstract
In this study, we investigate the aerodynamics of our newly designed golf ball that does not have dimples but grooves on
its surface. The smooth part of its surface is approximately 1.7 times that of a golf ball with dimples. We directly measure
the drag and lift forces on two versions of this golf ball in the ranges of real golf-ball velocity and rotational speed, and
compare them with those of smooth and dimpled balls. At zero spin, the drag coefficient of our balls shows a rapid fall-
off at a Reynolds number similar to that of a dimpled ball and maintains nearly a constant value (lower by 50% than that
of smooth ball) at higher Reynolds numbers. At non-zero spin, the lift-to-drag ratio of one version of our ball is higher
by 5%20% in the supercritical Reynolds number regime than that of a dimpled ball, but it is lower otherwise. With the
measured drag and lift forces, we predict the trajectories of our balls and compare them to those of smooth and
dimpled balls for the same initial conditions. The flight distances of our balls are larger by 148%202% than that of
smooth ball and shorter by 6%10% than that of a dimpled ball.

Keywords
Aerodynamics, golf ball, groove, lift, drag

Date received: 27 December 2013; accepted: 19 June 2014

Introduction that dimples cause local flow separations and trigger


the shear-layer instability along the separated shear
Increasing the flight distance has been one of the major layers, resulting in the generation of large turbulent
goals of golf-ball design. To achieve this goal, the sur- fluctuations, reattachment to the surface with high
face of a golf ball has been changed from smooth (e.g. momentum near the wall, and delay of main separa-
feathery golf ball in the 17th century) to dimpled sur- tion. This mechanism was confirmed by Smith et al.5
face, since a professor at Saint Andrews University dis- using direct numerical simulation. Nowadays, most
covered that a golf ball with a scored surface flew golf balls have about 250450 dimples on their surfaces
farther than that with smooth surface in mid-19th cen- to fully make use of their effect of reducing drag.
tury.1 Bearman and Harvey2 showed that dimples not Another goal of golf-ball design is to improve the
only reduce the drag on a golf ball by 50% but also putting accuracy because the putting accounts for
maintain the reduced drag coefficient, 38%50% of the total shots played in a round for both
CD = D=(0:5 rU20 A), nearly constant over a range of amateur and professional golfers.69 Jack Nicklaus, one
Reynolds numbers, Re = U0 d=n, where D is the drag, of the most accomplished professional golfers, noted
U0 is the free-stream velocity, d is the diameter of the the importance of putting in 1976, Undoubtedly, the
ball, r is the density, n is the kinematic viscosity, and majority of present-day tournaments are decided on
A (=pd2/4) is the cross-sectional area of the ball. Smits the putting surface. Recently, Alexander and Kern10
and Ogg3 considered an optimization of the dimple pat- examined the determinants of the earnings of
tern, including the dimple size and shape, to reduce
drag on a golf ball. They suggested that the dimple pat-
tern should be designed to promote turbulence at the Department of Mechanical & Aerospace Engineering, Seoul National
lowest possible Reynolds number without creating an University, Seoul, Korea
undesirable extra drag due to roughness. Choi et al.4
Corresponding author:
explained by measuring the streamwise velocity above Haecheon Choi, Department of Mechanical & Aerospace Engineering,
the dimpled surface how dimples on a sphere promote Seoul National University, Seoul 151744, Korea.
turbulence and lead to drag reduction. They showed Email: choi@snu.ac.kr

Downloaded from pip.sagepub.com at Monash University on December 6, 2014


2 Proc IMechE Part P: J Sports Engineering and Technology

Professional Golf Association (PGA) Tour golfers wind-tunnel experiments and compare them with those
from the period 19922001 and pointed out that the of smooth and dimpled balls. With the measured aero-
putting ability is the most important determinant of dynamic forces, the trajectories for the smooth,
earnings. dimpled, and G balls are computed for comparison.
Although they reduce the drag on the ball and
increase the flight distance, dimples are known to dete-
Methods
riorate the putting accuracy owing to their edges. As
Pelz8 pointed out, when the putter face contacts with Test balls
the edge of a dimple, the ball rotates and starts slightly Four golf-ball models are tested in this study: smooth,
off-line. To avoid this problem, one may draw a circle dimpled, and two G balls (Figure 1). These golf-ball
around the largest dimple-free area on a golf ball to models of 150 mm diameter are made of acrylonitrile
mark Dot-Spot.8 This spot becomes a hitting point butadiene styrene (ABS) resin. The present dimple-ball
so that a putter can make contact with the smooth sur- model (Figure 1(b)) is dynamically similar to a real golf
face of a golf ball. Despite the effort to enhance the ball (Titleist Pro V1x); 332 dimples are on the surface,
putting accuracy, only about 17% of putts from 4 m and their depth is about 0.9 mm. Figure 1(c) and (d)
were holed in the PGA tournaments during the years of shows two versions of G balls that are designed to be
19771992.8 This low rate of success is caused by the nearly spherically symmetric and thus may satisfy the
direction deviation in putting. Tierney and Coop11 United States Golf Association (USGA) rules. They
showed that the mean direction deviation in PGA Tour have 48 triangular and 6 tetragonal smooth surfaces
veterans was 1.3% of the putting distance. Karlsen12 surrounded by net-shaped grooves, and the depth and
reported that the mean putting deviation during repeti- the width of grooves in G balls are 0.8 and 10 mm,
tive putts for eight elite players was 1.8% on a flat respectively. The net-shaped grooves are connected to
indoor green. Recently, Caesar Golf Company made a (Figure 1(c); denoted as G1 ball hereafter) or separated
perfectly smooth golf ball, The Caesar Featherie, by from others (Figure 1(d); denoted as G2 ball hereafter).
eliminating dimples on the surface. In the review of this The G2 ball has more smooth surfaces than the G1 ball
ball, however, it was expected that the Caesar Featherie because pentagon- or hexagon-shaped smooth parts
would drop quickly and achieve anywhere from 30% to (gray color in Figure 1(d)) are added to the G1 ball at
80% of a clubs normal distance because this ball is all the intersection points of the grooves.
designed without dimples for accuracy rather than for To show the superior putting accuracy of G balls, we
distance.13 Therefore, it is very desirable to design a drop two balls (dimpled and G1 balls, respectively) on a
new golf ball that has both high putting accuracy and flat surface and record their bounce trajectories using a
good aerodynamic performance (i.e. large flight high-speed camera (Photron FASTCAM Super 10K) at
distance). 250 fps. As shown in Figure 2, the G1 ball bounces back
Recently, we devised a new golf ball that did not toward its original position when it is dropped with its
have dimples but grooves on its surface.14,15 This ball is flat area facing down in the direction of gravity. On the
much smoother than conventional dimpled balls. For other hand, the dimpled ball bounces back with a devia-
example, the area of grooves is less than 50% of total tion of 13 from its original position. This amount of
surface, whereas that of dimples is larger than 70%. deviation angle agrees well with the result of Pelz.8
Therefore, accurate putts are naturally expected for the
ball by Choi et al.14 (called G ball hereafter). Since
the G ball has higher putting accuracy than dimpled Wind-tunnel experiments
balls (see below), an important question is now how far Figure 3 shows the schematic diagram of the present
this ball flies. To investigate the aerodynamic perfor- experimental setup, consisting of an open-type wind
mance of the G ball, we directly measure the drag and tunnel, golf-ball model, motor, supporter, and load cell.
lift forces on the two versions of G balls in the ranges The same wind tunnel is used as that described by Jeon
of real golf-ball velocity and rotational speed through et al.,16 and the background turbulence intensity at the

Figure 1. Four golf-ball models tested: (a) smooth ball, (b) dimpled ball, (c) G1 ball, and (d) G2 ball.

Downloaded from pip.sagepub.com at Monash University on December 6, 2014


Kim and Choi 3

Figure 3. Schematic diagram of the experimental setup.

enough such that the side walls have little effect on the
flow in the wake and the drag and lift forces. A small
direct current (DC) motor is installed inside the golf-
ball model and rotates it about the vertical axis. The
vibration of the model due to the rotation is minimized
by avoiding the resonance of the system. The rotational
speed of the model is varied by adjusting the voltage
applied to the motor and measured using a tachometer.
The model with motor is fixed to the supporter that is
designed to minimize the support interference. The
upper part of the supporter is a pipe in which two wires
supply a voltage to the motor inside the model. The dia-
meter of this pipe is only 2% of the golf-ball model dia-
Figure 2. Bounce trajectories of the (a) dimpled ball and (b) meter. The lower part of the supporter is shaped into an
G1 ball. The numbers (19) in this figure denote the time
ellipse with a ratio of major to minor axis of 2. The drag
sequence and the consecutive time interval is 0.02 s.
and lift on the model are measured using a load cell
(CAS BCL-3L) attached to the supporter. The calibra-
free-stream velocity of 10 m/s is lower than 0.5%. Son tion curve is linear, and the uncertainty of force mea-
et al.17 showed that the background turbulence inten- surement is 62.5%. The output from the load cell is
sity within 0.5% has little effect on the flow over a amplified and sampled for 60 s at the rate of 32 kHz to
sphere at the Reynolds number range investigated in obtain a fully converged mean drag and lift. The data
this study. The cross section of the wind tunnel after from the measurement are transferred to a computer
contraction is 600 mm 3 600 mm, and the maximum through an analog-to-digital (A/D) converter (NI PCI-
wind speed at the test section is 30 m/s. The golf-ball 6251) and post-processed. Note that, in this study, the
model is installed at the center of the test section, and golf-ball model is supported from the bottom for its
the side walls are located at two times the diameter of rotation. Kray et al.20 showed that the drag coefficient
the golf-ball model away from its center. The blockage of the sphere supported from the bottom is higher than
ratio of the cross-sectional area of the golf-ball model that supported from the rear due to the support inter-
to the test-sectional area is 4.9%, which is below the ference in the first case. In their study, the supporter on
critical value ensuring negligible blockage effect accord- the bottom was rotated together with a sphere and the
ing to Achenbach.18 Stanczak et al.19 measured the drag on the rotating supporter was not directly mea-
wake behind rotating dimpled balls using a hot-wire sured but assessed from the previous results21 of a
anemometry and showed that the size of tilted wake at rotating cylinder. On the other hand, in this study, the
x/d \ 2 is smaller than the ball diameter, where x is the supporter does not rotate and the measured forces are
streamwise distance from the center of the dimpled ball. corrected by subtracting those of isolated supporter
Therefore, the test section of our wind tunnel is wide measured separately. This force measurement method

Downloaded from pip.sagepub.com at Monash University on December 6, 2014


4 Proc IMechE Part P: J Sports Engineering and Technology

has been used by our group in the past22,23 and shown


to be accurate (see below).
The free-stream velocity U0 varies from 5 to 27 m/s,
corresponding to the Reynolds numbers,
Re = 0.5 3 1052.7 3 105. The rotational speed (N) of
the golf-ball model varies from 0 (stationary) to
330 r/min, corresponding to the spin ratios (ratio of
maximum surface velocity to the free-stream velocity),
a = pdN/(60U0) = 00.52. These cover the ranges of
real golf-ball Reynolds number and spin ratio.24 The
drag and lift coefficients, CD and CL, are defined as Figure 4. Trajectory computation.

D L
CD = , CL = 1
0:5 rU20 A 0:5 rU20 A Table 1. Initial conditions for average driver shots2628 (see
Figure 4).

where L is the lift on the golf-ball model. v0 (m/s) u0 () v0 (rad/s)

PGA tour 74.7 10.9 281.3


Trajectory computation LPGA tour 62.1 13.2 273.4
Male amateur with a 61.7 11.9 334.3
With the measured aerodynamic forces and initial con- handicap of 10
ditions, it may be possible to compute the trajectory of Female amateur with a 53.2 12.4 335.8
handicap of 10
a golf-ball drive. For the computation of trajectories,
we assume that the golf ball satisfies a quasi-steady PGA: Professional Golf Association; LPGA: Ladies Professional Golf
state at every instance. The drag and lift coefficients Association.
are obtained from the interpolation of measured data.
The equations of motion2 derived from the Newtons
second law are
"   2 #
d2 x rA dx 2 dy
= + (CD cos u + CL sin u)
dt2 2m dt dt
2

"   2 #
d2 y rA dx 2 dy
2
= + (CL cos u  CD sin u)  g
dt 2m dt dt
3

where x and y are the positions of the ball in the hori-


zontal and vertical directions, respectively; m is the
mass of the ball; u is the inclination angle of the flight Figure 5. Variation in the drag coefficient with the Reynolds
path to the horizontal line; and g is the gravitational number for stationary (non-rotating) balls: s, smooth ball29; ,
acceleration (Figure 4). The effect of spin decay during smooth ball (present); 3, dimpled ball2;M, dimpled ball4; N,
dimpled ball (present); h, G1 ball (present); and j, G2 ball
a flight is also considered using the exponential spin-
(present).
decay model proposed by Smits and Smith25
 
2cv
v(t) = v0 exp  t 4 Results and discussion
d
Drag variation on stationary golf balls
where v is the rotational speed (rad/s), v0 is the initial Figure 5 shows the drag coefficients of smooth,
rotational speed, c is the experimental constant dimpled, and two G balls, together with those of previ-
(=5 3 1025), v is the velocity, and t is the time. Initial ous studies. The present drag coefficient of a smooth
conditions are determined on the basis of the average ball shows good agreement with that of Achenbach.29
professional and amateur driver shots as shown in Also, the present result on the dimpled ball shows that
Table 1.2628 A fourth-order RungeKutta (RK4) the minimum drag coefficient is between the results of
method is used to integrate equations (2) and (3) with Bearman and Harvey2 and Choi et al.4 This is because
the computational time step size of Dt = 0.001 s. the depth of present dimples (k/d 0.6 3 1022) is

Downloaded from pip.sagepub.com at Monash University on December 6, 2014


Kim and Choi 5

Figure 6. Variations in the (a) drag coefficient, (b) lift coefficient, and (c) lift-to-drag ratio for dimpled (left) and G2 (right) balls with
the Reynolds number at N = 0330 r/min: s, N = 0 r/min (no spin); N, 150 r/min; j, 250 r/min; and , 330 r/min.

between those of Bearman and Harvey2 (k/d = 0.9 3 similar to that of the present dimpled ball, showing
1022) and Choi et al.4 (k/d = 0.4 3 1022), where k is about 50% drag reduction as compared to that of
the dimple depth. These good agreements in the drag smooth ball. For both the dimpled and G2 balls, the
coefficients of smooth and dimpled balls indicate the reduced drag coefficients remain nearly constant at
reliability of the present results in the subcritical, criti- supercritical Reynolds numbers of Re = 1.1 3 105
cal, and supercritical Reynolds number ranges. 2.7 3 105.
Like the dimpled ball, the drag coefficients of two G It is interesting to note that the drag coefficient of
balls start to decrease at much lower Reynolds numbers the G2 ball decreases faster than that of the G1 ball,
than that of the smooth ball, indicating that the multi- although the G2 ball has more smooth surfaces. The
ple grooves on the surfaces of G balls result in an aero- role of grooves in drag reduction is to trigger the flow
dynamic characteristics similar to that of dimples. The instability along the shear layer separated from the
G2 ball shows a rapid fall-off of CD at a Reynolds edges of the grooves, and this mechanism of grooves is
number similar to that of the dimpled ball. The drag similar to that of dimples.4,5 As mentioned previously,
coefficient of the G2 ball is slightly higher near the crit- in the G2 ball, pentagon- or hexagon-shaped smooth
ical Reynolds number range than that of the present parts (gray color in Figure 1(d)) are added to the G1
dimpled ball, but the minimum drag coefficient is ball at all the intersection points of the grooves for the

Downloaded from pip.sagepub.com at Monash University on December 6, 2014


6 Proc IMechE Part P: J Sports Engineering and Technology

G2 balls with the Reynolds number at N = 150


330 r/min, together with those of stationary (0 r/min)
balls. At supercritical Reynolds numbers of
Re = 1.1 3 1052.7 3 105, the drag coefficients of both
balls decrease with increasing Reynolds number at a
given rotational speed, but they increase with increas-
ing rotational speed (Figure 6(a)). The present result
agrees well with that of Bearman and Harvey.2 The lift
coefficients show the same behavior as the drag coeffi-
cients (Figure 6(b)). The behavior of the lift-to-drag
ratio is also similar to that of the lift coefficient, in that
it increases with the rotational speed and decreases with
increasing Reynolds number in the supercritical regime.
In the critical regime, it is interesting to note that the
lift coefficient becomes negative at Re = 0.5 3 105 and
N = 150 r/min (Figure 6(b)). This phenomenon is
closely related to the asymmetric boundary-layer
separation, which is discussed below.
The variations in the drag and lift coefficients with
the spin ratio are shown in Figure 7(a) and (b). At
Re = 1.1 3 1052.7 3 105 (supercritical Reynolds num-
bers), the drag and lift coefficients measured from dif-
ferent Reynolds numbers collapse well into one line
and increase with the spin ratio. This behavior agrees
well with the result of Bearman and Harvey2 investi-
gated for a dimpled ball. On the retreating side, the ball
surface moves in the same direction with the flow and
the near-wall momentum increases with the spin ratio,
resulting in the separation delay. Thus, the lift coeffi-
cient increases with the spin ratio. In the supercritical
Reynolds numbers, the drag coefficients of the G2 ball
are higher by 4%10% than those of the dimpled ball
(e.g. at Re = 1.9 3 105 and a = 0.14, the drag coeffi-
cients of the dimpled and G2 balls are 0.28 and 0.30,
respectively), but its lift coefficients are also higher by
10%27% than those of the dimpled ball (e.g. at
Re = 1.9 3 105 and a = 0.14, the lift coefficients of the
dimpled and G2 balls are 0.20 and 0.25, respectively).
On the other hand, at Re = 0.5 3 1050.9 3 105 (crit-
ical Reynolds numbers), the drag and lift coefficients
Figure 7. Variations in the (a) drag coefficient, (b) lift deviate from those at supercritical Reynolds numbers
coefficient, and (c) lift-to-drag ratio for dimpled (black) and G2 and show more complicated behaviors. This is because

(red) balls with the spin ratio: , Re = 0.5 3 105; , 0.7 3 105; s,
laminar separation occurs on the retreating side and
0.9 3 105; h, 1.1 3 105; M, 1.3 3 105; P, 1.5 3 105; , 1.7 3 105; turbulent separation on the advancing side (i.e. side
j, 1.9 3 105; N, 2.1 3 105; H, 2.3 3 105; + , 2.5 3 105; and 3, moving against the flow) at critical Reynolds numbers,
2.7 3 105. whereas turbulent separations occur on both sides at
supercritical Reynolds numbers.30 The asymmetric
purpose of blocking possible flow paths inside the boundary-layer separation was also observed for a
grooves. From the edges on the smooth parts at the cricket ball because the seam trips the laminar bound-
intersection points as well as those of grooves, the ary layer into turbulence on only one side of the ball.3133
shear-layer instability occurs and thus the flow separa- As mentioned before, the lift coefficient of the dimpled
tion is delayed, resulting in more drag reduction from ball becomes negative at Re = 0.5 3 105 and a = 0.24.
the G2 ball. This phenomenon is called the inverse (or negative)
Magnus effect. The negative lift on a rotating golf ball
with dimples was also observed experimentally and
Drag, lift, and lift-to-drag ratio of rotating golf balls numerically.2,30 The lift coefficient of the G2 ball also
Figure 6 shows the variations in the drag and lift coeffi- becomes negative at Re = 0.5 3 105 and a = 0.24, but is
cients and lift-to-drag ratio of the rotating dimpled and close to 0.

Downloaded from pip.sagepub.com at Monash University on December 6, 2014


Kim and Choi 7

Figure 8. Computed trajectories and the corresponding flight distances (in meter) of the smooth, dimpled, G1, and G2 balls for
the same initial conditions of (a) PGA Tour, (b) LPGA Tour, (c) male amateur with a handicap of 10, and (d) female amateur with a
handicap of 10: .., smooth; - - -, dimpled; -  -, G1; and , G2.

Figure 7(c) shows the lift-to-drag ratio (i.e. the aero- smooth, dimpled, G1, and G2 balls for the same initial
dynamic performance) for the dimpled and G2 balls. conditions (Table 1). The flight distance of the dimpled
At Re = 1.1 3 1052.7 3 105, the lift-to-drag ratios of ball is about 1.72.2 times larger than that of the
the G2 ball are higher by 5%20% than those of the smooth ball. For the average PGA and Ladies
dimpled ball (e.g. at Re = 1.9 3 105 and a = 0.14, the Professional Golf Association (LPGA) Tour players
lift-to-drag ratios of the dimpled and G2 balls are 0.71 (Figure 8(a) and (b)), the flight distances of G1 and G2
and 0.83, respectively), implying that the aerodynamic balls are about 90%91% and 94% that of the dimpled
performance of the G2 ball is better at supercritical ball, respectively. As shown in Figure 8(c) and (d), sim-
Reynolds numbers and spin ratios considered. On the ilar trends are observed for the average male and
other hand, at Re = 0.5 3 1050.9 3 105, the perfor- female amateur players.
mance of the G2 ball is worse than that of the dimpled The Reynolds numbers and lift-to-drag ratios of the
ball. As discussed in the below, the flight trajectory of dimpled and G2 balls during the flight are shown in
each ball is influenced by this L/D characteristics at the Figure 9 in the case of PGA Tour average driver shot.
critical and supercritical Reynolds numbers. The Reynolds numbers of both balls decrease as they
fly and show slight increases at the ends of the flights
owing to the gravitational acceleration. In the early
Flight trajectory period of the flight (supercritical Reynolds number
Figure 8(a)(d) shows the computed trajectories and range), L/D of the G2 ball is larger by up to 18% than
the corresponding flight distances (in meter) of the that of the dimpled ball (e.g. at t = 0.1 s, the lift-to-drag

Downloaded from pip.sagepub.com at Monash University on December 6, 2014


8 Proc IMechE Part P: J Sports Engineering and Technology

gravitational acceleration (Figure 9), but the rotational


speeds keep decreasing, and thus, the spin ratios reduce
rapidly. This behavior agrees well with the result of
Beasley and Camp24 investigated for a dimpled ball.

Conclusion
In this study, we evaluated the aerodynamic perfor-
mances of our newly designed golf balls with grooves on
its surface (called G1 and G2 balls) through wind-tunnel
experiments. At zero spin (a = 0), the drag coefficient of
G2 ball showed a rapid fall-off at a critical Reynolds
number and maintained a constant value (50% lower
Figure 9. Variations in the flight Reynolds number and lift-to- than that on smooth ball) at larger Reynolds numbers,
drag ratio in time for the dimpled and G2 balls in the case of indicating that the multiple grooves on the surface
PGA Tour average driver shot: - - -, Re (dimpled); , Re (G2); resulted in a similar aerodynamic characteristics to those
s, L/D (dimpled); and M, L/D (G2). of dimples at zero spin. At non-zero a (i.e. rotating ball),
both the drag and lift coefficients of the G2 ball were
higher than those of the dimpled ball in the supercritical
Reynolds number regime, but its lift-to-drag ratio was
also higher by about 5%20% than that of the dimpled
ball. However, at the critical Reynolds number regime,
the aerodynamic performance of the G2 ball was worse
than that of the dimpled ball.
With the measured aerodynamic forces, the trajec-
tories of the smooth, dimpled, G1, and G2 balls were
computed for same initial condition. In the early period
of the flight, the lift-to-drag ratio of G2 ball was higher
than that of the dimpled ball, and thus, the G2 ball flied
higher than the dimpled ball. As the flight Reynolds
number of G2 ball changed from the supercritical to
critical flow regime, however, the lift-to-drag ratio of
G2 ball became much lower than that of the dimpled
Figure 10. Variations in the flight Reynolds number and spin ball and the G2 ball rapidly descended. As a result, for
ratio in time for the dimpled and G2 balls in the case of PGA the average professional and amateur players, the flight
Tour average driver shot: - - -, dimpled and , G2. distances of G2 ball were about 1.62 times larger than
those of smooth ball, but was 4.5%6% shorter than
ratios of the dimpled and G2 balls are 0.59 and 0.70, those of the dimpled ball, respectively. Note that the
respectively), and thus the G2 ball flies higher than the present flight distances were obtained with the quasi-
dimpled ball as shown in Figure 8(a). However, as the steady state assumption and spin-decay model in equa-
speed of the ball decreases owing to the drag, the flight tions (2)(4).
Reynolds number decreases and reaches the critical Although the flight distances of G balls are slightly
Reynolds number regime. Then, the G2 ball rapidly shorter than that of the dimpled ball, the G balls have
descends as seen in Figure 8(a) because of lower lift-to- more smooth surfaces and provide more accurate put-
drag ratios in the critical Reynolds number regime ting. For better aerodynamic performance of the G
(Figure 9), resulting in a shorter flight distance than balls, a strategy of increasing the lift-to-drag ratio in
that of the dimpled ball. the critical Reynolds number range should be devel-
Figure 10 shows the variations in the flight Reynolds oped, which is underway in our group.
number and spin ratio in time for the dimpled and G2
balls in the case of PGA tour average driver shot Declaration of conflicting interests
(Table 1). As mentioned previously, the spin decay dur- The authors declare that there is no conflict of interest.
ing a flight is considered using the model, equation (4),
proposed by Smits and Smith.25 The spin ratios of both
balls gradually increase as they fly, because the speeds Funding
of both balls decrease faster than the rotational speeds This work was supported by the research programs
do. However, at the end of the flight, the flight (2011-0028032, 2014M3C1B1033980) of NRF, MSIP,
Reynolds numbers increase with time owing to the Korea.

Downloaded from pip.sagepub.com at Monash University on December 6, 2014


Kim and Choi 9

References 18. Achenbach E. The effects of surface roughness and tunnel


1. Chase A. A slice of golf. Science 1981; 81: 9091. blockage on the flow past spheres. J Fluid Mech 1974; 65:
2. Bearman PW and Harvey JK. Golf ball aerodynamics. 113125.
Aeronaut Quart 1976; 27: 112122. 19. Stanczak MB, Lemons LD, Beasley DE, et al. Observa-
3. Smits AJ and Ogg S. Aerodynamics of the golf ball. In: tions of the wake characteristics of spinning golf balls.
Hung GK and Pallis JM (eds) Biomedical engineering In: The third world scientific congress of golf, St Andrews,
principles in sports. New York: Springer, 2004, pp.327. 2024 July 1998, pp.440456. Champaign, IL: Human
4. Choi J, Jeon W-P and Choi H. Mechanism of drag Kinetics.
reduction by dimples on a sphere. Phys Fluids 2006; 18: 20. Kray T, Franke J and Frank W. Magnus effect on a
041702. rotating sphere at high Reynolds numbers. J Wind Eng
5. Smith C, Beratlis N, Balaras E, et al. Numerical investi- Ind Aerod 2012; 110: 19.
gation of the flow over a golf ball in the subcritical and 21. Swanson WM. The Magnus effect: a summary of investi-
supercritical regimes. Int J Heat Fluid Fl 2010; 31: 262 gations to date. J Basic Eng: T ASME 1961; 83: 461470.
273. 22. Park H and Choi H. Aerodynamic characteristics of fly-
6. Palmer A and Dobereiner P. Arnold Palmers complete ing fish in gliding flight. J Exp Biol 2010; 213: 32693279.
book of putting. London: Stanley Paul, 1986. 23. Park H, Bae K, Lee B, et al. Aerodynamic performance
7. Wiren G. Golf: building a solid game. Englewood Cliffs, of a gliding swallowtail butterfly wing model. Exp Mech
NJ: Prentice Hall, 1987. 2010; 50: 13131321.
8. Pelz D. Dave Pelzs putting bible: the complete guide to 24. Beasley D and Camp T. Effects of dimple design on the
mastering the green. New York: Random House Digital, aerodynamic performance of a golf ball. In: The fourth
Inc., 2000. world scientific congress of golf, St Andrews, 2226 July
9. Dorling K. Golf skills. London: Dorling Kindersley Ltd, 2002, pp.328340. London: Routledge.
2011. 25. Smits A and Smith D. A new aerodynamic model of a
10. Alexander DL and Kern W. Drive for show and putt for golf ball in flight. In: The second world scientific congress
dough? An analysis of the earnings of PGA Tour golfers. of golf, St Andrews, 48 July 1994, pp.340347. London:
J Sport Econ 2005; 6: 4660. E & FN Spon.
11. Tierney DE and Coop RH. A bivariate probability model 26. TrackMan company. Ball speed, http://mytrackman.
for putting proficiency. In: The third world scientific con- com/explore/trackman-data/trackman-ball-data/ball-speed
gress of golf, St Andrews, 2024 July 1998, pp.385394. (accessed 10 April 2014).
Champaign, IL: Human Kinetics. 27. TrackMan company. Launch angle, http://mytrackman.
12. Karlsen J. Golf putting: an analysis of elite-players tech- com/explore/trackman-data/trackman-ball-data/launch-
nique and performance. Masters Thesis, Norwegian Uni- angle (accessed 10 April 2014).
versity of Sport and Physical Education, Oslo, 2003. 28. TrackMan company. Spin rate, http://mytrackman.com/
13. Barzeski EJ. Caesar Featherie dimpleless golf ball review. explore/trackman-data/trackman-ball-data/spin-rate (accessed
TheSandTrap.com, http://thesandtrap.com/b/balls/cae- 10 April 2014).
sar_featherie_dimpleless_golf_ball_review (2008, accessed 29. Achenbach E. Experiments on the flow past spheres at
26 December 2013). very high Reynolds numbers. J Fluid Mech 1972; 54:
14. Choi HC, Choi J and Son GM. Golf ball. Patent 565575.
EP2112948, England, Germany, and France, 2013. 30. Beratlis N, Squires K and Balaras E. Numerical investiga-
15. Kim J and Choi H. Aerodynamics of a golf ball with tion of Magnus effect on dimpled spheres. J Turbul 2012;
grooves. In: Proceedings of the seventh international sym- 13: 115.
posium on turbulence and shear flow phenomena, Ottawa, 31. Mehta RD. Aerodynamics of sports balls. Annu Rev Fluid
ON, Canada, 2831 July 2011, paper no. 4D2P. Avail- Mech 1985; 17: 151189.
able at: http://www.tsfp-conference.org/images/stories/ 32. Lock GD, Edwards S and Almond DP. Flow visualiza-
proceedings/2011/4d2p.pdf tion experiments demonstrating the reverse swing of a
16. Jeon S, Choi J, Jeon W-P, et al. Active control of flow cricket ball. Proc IMechE, Part P: J Sports Engineering
over a sphere for drag reduction at a subcritical Reynolds and Technology 2010; 224: 191199.
number. J Fluid Mech 2004; 517: 113129. 33. Scobie JA, Pickering SG, Almond DP, et al. Fluid
17. Son K, Choi J, Jeon W-P, et al. Effect of free-stream tur- dynamics of cricket ball swing. Proc IMechE, Part P: J
bulence on the flow over a sphere. Phys Fluids 2010; 22: Sports Engineering and Technology 2013; 227: 196208.
045101.

Downloaded from pip.sagepub.com at Monash University on December 6, 2014

S-ar putea să vă placă și