Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

1

I N T R O DU C T IO N

Golf is unique among games for the sheer perfection of its range of equipment. In the
wider category of sports, we might compare the perfection of golf clubs and golf balls
to the bows and arrows in modern archery, or even to the range of high-technology
bicycles used for the different stages of the Tour de France. The changes in golf
equipment are less obvious, mainly due the work of the governing bodies of golf in
keeping the game as close as possible to its traditional form. However, the science
underlying these changes, including ones from the beginning of the 20th century,
are equally profound. Frank Thomas titled his book, on the evolution of the golf
equipment rules, From Sticks and Stones (Frankly Golf Publications, 2011). The title
was clearly meant to imply an evolution from the simple clubs and balls used at the
time of the first code of golf rules, established in 1744 by the Honourable Company
of Edinburgh Golfers. However, the title is much more prescient than it appears.
The skills exhibited in golf are based on our unique physical ability to maintain
a highly stable upright stance while the arms and upper body perform powerful ath-
letic tasks. This allowed our earliest ancestors to throw stones and swing sticks with
great precision; and later to wield hammers, axes, and even golf clubs with precise
purpose. Children as young as four have an innate ability to perform these tasks.
Some with exceptional athletic ability can drive a golf ball after watching a skilled
golfer perform the task. Most can throw balls quite accurately underarm; by the age
of six, without instruction, they naturally apply a stiff-arm swing, thereby reducing
the number of degrees of freedom for increased repeatability when asked to throw
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

underarm at targets (Jacques et al. 1989). A precisely repeatable swing is of course


of little value without the stability of the stance; in fact, none of the primates with
much greater strength and agility, but without a stable upright stance, can accom-
plish any of these tasks. This is the essence of the golf swing and the putting swing;
that is, maintain the greatest stability to keep the rotation “center” of the swing as
stationary as possible. This issue will be addressed in Chapter 3.
Before embarking on the story of golf science, we will consider some compari-
sons of performance from the car and aerospace world to put golf performance in
some perspective. We will introduce some science to do this, which is not difficult to
follow. All of the book chapters have easy-to-follow explanations presented in a sim-
ilar form. So if you follow these initial arguments, you will be able to follow the sci-
ence of all aspects of the game from the club swing, through the impact-generated
launch conditions into ball flight, to the final roll of the ball on the green. The sci-
ence story of the game is truly remarkable.
The back of each chapter contains the full “Details of the Modeling,” which necessi-
tates some heavier physics. It is kept to the minimum possible but is needed to provide

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.
2 Introduction

150 mph 1,500 lb

15 seconds

Figure 1A  Average force from back wheels to go from 0 to 150 mph in 15 seconds.

the necessary support for some of the very surprising conclusions that are reached
throughout the text. These sections comprise about one-third of the total text and can
be skipped without any loss of understanding of the science principles of the game.
For others with a deeper science-oriented background, the physics sections will
hopefully stimulate discussion, and in some cases, further investigations.
Let’s start with a fantasy trip, to explain one of the most important science
aspects of golf. A 15-second period of this trip is illustrated in Figure 1A. Assume
you are heading out to the golf course in your Ferrari 730 hp F12berlinetta. Starting
through an intersection is an empty stretch of open road, leading to the course. You
step on the gas and go from zero to 150 miles per hour in 15 seconds; and coming
up over a rise and around a curve, you do a 2 g deceleration, which with much tire
squealing brings you back down to 30 miles per hour in 120 yards for a smooth turn
into the golf club. Don’t worry if that’s just a dream, because you are about to do
something even more amazing on the first tee. To appreciate this fact, we will use
the car ride to understand the principles involved:

Recall that Newton told us that


force equals mass times acceleration,

and during the car sprint, your average acceleration was 10 miles per hour per
second. So we can form a second relationship. The final speed is the acceleration
multiplied by the time, so we can multiply our relationship by time and get
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

force × time = mass × acceleration × time = mass × final speed.

Now it’s just a matter of using the language of impact:

force × time = impulse; mass × final speed = momentum.

More generally, impulse is always equal to change of momentum.


The mass of our Ferrari is 1,500 kilograms (3,300 pounds), and the final speed
is 67 meters per second (150 miles per hour). The conversions to metric units are
given to simplify the calculations. The problem here is mass, and strictly speaking
weight is not mass. Physicists get concerned about the difference because weight is
the force of gravity on an object that deflects the weighing scale. In space, an object
can be weightless, but it still retains its mass. Since all golf takes place on the surface
of the earth, it’s actually OK to talk about a driver head “mass” of 0.44 pounds, and a
ball “mass” of 0.10 pounds. But, if we use pounds with Newton’s laws, then we have

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.
3 Introduction

150 mph 1,500 lb

0.00045 seconds

Figure 1B  Average force from the club face impact for the ball to accelerate from 0 to 150 mph in
0.00045 seconds.

to cancel out gravity, which makes for messy calculations. If we work with kilograms
and meters per second, then we just plug the values straight in. So we have “force ×
15 seconds” equals “1,500 kilograms × 67 meters per second”; or force = (1,500 ×
67)/15, which gives 6,700 newtons of force. One pound is equal to approximately
4.5 newtons, and if you remember the tale about Newton putting all this together by
watching an apple fall from a tree, then it makes sense because we typically get about
4 or 5 apples in a pound. Therefore the average force from the back wheels of your
Ferrari, accelerating it forward, is 6,700/4.5 = 1,500 pounds.
Now you are on the first tee. It’s going to be a great morning, you just hit one
of your best drives, and with some bounce and roll, it looks like it’s out there about
250 yards. To do this you must have hit the ball around 100 miles per hour, and it
took off at approximately 150 miles per hour. For comparison with your Ferrari,
it did this, from a standing start, as shown in Figure 1B, in just 0.00045 seconds!
Surprisingly, the force you applied to the ball was just about the same as the driv-
ing force produced by the wheels of your 730 hp Ferrari. In this case, we have a
ball mass of 0.045 kilograms (0.1 pounds) instead of 1,500 kilograms and a time
of 0.00045 seconds instead of 15 seconds. So, the calculation for force changes to
“force × 0.00045” equals “0.045 × 67”; or force = (0.045 × 67)/0.00045, which gives
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

6,700 newtons or 1,500 pounds, exactly as before. In this case, the force starts from
zero, reaches a maximum at just over 0.0002 seconds, and then decreases back to
zero at 0.00045 seconds as the ball leaves contact. The maximum force is thus close
to 3,000 pounds; for a better feeling of this magnitude, let’s say a “ton and a half.” So
just laugh when you see those four extra miles per hour claims for “low friction” tees,
which magically add 80 pounds to the maximum impact force.
We can break contact time down into two parts. At just over 0.0002 seconds,
the ball was compressed to about a 1 inch diameter imprint on the face, and the
club head had slowed down to about 82 miles per hour, with the ball travelling
with it at the same speed. If you had hit a trick sticky golf ball filled with thick
molasses, instead of a real golf ball with a rubber core, the process would end at
this point, and the compressed ball would remain compressed and sticking to the
face. At this point, the mass at the end of the shaft would be both the 0.44 pound
head and the 0.1 pound ball. From Newton’s laws it can be shown that, slowed
down to 82 miles per hour, the energy of the moving mass at the end of the shaft
would still be 130 foot-pounds. As you get swept off your feet trying to slow down

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.
4 Introduction

this handful of energy, be consoled that you have just experienced zero coefficient
of restitution.
Instead, a real golf ball starts to spring off the face shortly after 0.0002 seconds;
the action of restitution or recovery speeds the ball up to 150 miles per hour, while
the reaction slows the club head down further to 67 miles per hour. Therefore, com-
pared to the 100 miles per hour strike against a stationary ball, at the end of the
impact we have a 150 miles per hour ball and a 67 miles per hour club head with a
speed difference between the two of 83 miles per hour. So the golf ball, compared to
the molasses-filled ball, has recovered 83 percent of the impact speed. In proportion
terms, 0.83 of the impact speed has been recovered in speed away from the moving
club face; of course, this is the coefficient of the restitution for the impact, abbrevi-
ated throughout the book as CofR. The energy of the 0.44 pound head travelling at
67 miles per hour has been halved during restitution to 65 foot-pounds, which in
slowing down just provides a smooth follow-through, finishing in an elegant stance
with the belt buckle facing the target—nicely done!
However, the beautiful swing would have accomplished very little without the
dimples on the surface of the ball, first introduced in the early 1900s and now per-
fected in shape through exhaustive computer modeling and wind-tunnel testing.
Using a smooth round ball, even with 0.83 CofR, your drive would probably have
carried about 130 yards, with maybe bounce and roll taking it another 20 yards. We
will wait until the next shot to figure that one out.
The bounce and roll must be better than you thought because you are now only
115 yards out—just made for your pitching wedge shot. Your magic golf day con-
tinues. The ball launches around 40 degrees and just seems to keep on climbing, up
to somewhere around 80 feet. From its steep descent, it takes a single short bounce
about 10 feet from the hole, follows a smooth spiral trajectory as it rolls off the sloped
right side of the green, and has slowed down to about 1 mile per hour before crossing
just inside the edge of the hole. Any faster and it would have escaped, but instead it
rolls half way around the edge of the hole without touching the flag stick, runs out of
steam, and topples in sideways—maybe time to just go and celebrate.
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

The eagle approach shot happened because you managed to put around
8,000 revolutions per minute backspin on the ball, and in so doing changed it to a
little Harrier jump jet, which just kept on climbing as it appeared to do. The story of
this propulsion system, with its associated low-drag performance, must wait until
Chapter 2. The ball actually landed with most of this backspin still in place, which
caused it to check quickly and start its slow forward roll.
One final comparison with the Ferrari, and then we will move onward. The
amazing amount of backspin is produced because the ball sticks to the club face,
exactly like the rubber car wheels stuck to the road in the braking turn. If we could
put a microphone on the wedge, then amplify and slow down the signal, we would
likely detect very high-frequency squealing as the ball is forced to rotate while grip-
ping the face. This is exactly the same as the Ferrari tires gripping the road around
the bend while different points on the area of contact must travel at different speeds.
An even better comparison is the squealing of rubber soled athletic shoes as they
brake and accelerate during the step, all the while gripping the floor. Not coinciden-
tally, the latest high-spin golf balls have covers made from the same urethane poly-
mers as indoor running shoes. We will see later in the book how this produces such

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.
5 Introduction

amazing amounts of backspin that the ball actually comes off the face skidding, just
as if a powerful internal spring had been wound up and then released—which is in
fact what actually happens.
It’s time to describe some of the important equipment developments that have
taken the game “from stick and stones” to the exciting game of today. We will
restrict the discussion to approximately the last half-century. The interested reader
may wish to consult Thomas (2008 and 2011) for a more detailed discussion of this
topic and an enjoyable discussion of the early history of the game.

I MPORTA NT DEV ELOPM ENTS IN GOLF


EQU IPM ENT OV ER THE PAST 50 Y E A R S
A primary source for understanding the fundamental aspects of golf is still the super-
lative study, titled Search for the Perfect Swing, by Cochran and Stobbs ([1968] 1999).
This study, which started in the 1960s and involved the collaboration of academics
from a number of research institutions in Britain, resulted in a wealth of information
that, 50 years on, still forms a solid foundation for the scientific study of the game.
The title of the present work is intended, in part, to pay tribute to many of the expla-
nations in Cochran and Stobbs, which were made before the existence of invaluable
experimental data from wind-tunnel testing, ultra high-speed photography, laser
monitoring, and Doppler radar tracking systems. However, the title has more to do
with the performance of PGA and LPGA Tour players, which has been captured by
the Trackman Company, and which over the aggregate represent as near to perfect
golf as it is possible to measure. The overall goal of this text has been to lay down, as
well as possible, the science necessary to describe the ball striking of this group of
elite players, and at the same time to try to establish some measure of the improve-
ment in performance of golf equipment in recent time, particularly for the modern
titanium driver and the current generation of premium golf balls.
Since the time of the “perfect swing” study, according to Thomas (Just Hit It,
2008), the Technical Director of the USGA from 1974 to 2000, only three equip-
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

ment innovations have made a profound change to the game of golf. These are
perimeter weighting of club heads, the graphite shaft, and the incorporation of
spring effect into the face of hollow titanium drivers. To this list should be added
the latest premium golf balls, designed to perform quite differently with the driver
and the grooved irons. But even this can be attributed to the thin-face titanium
driver, which from the necessity of surviving multiple impacts could not have a
grooved face.
The use of perimeter weighting, producing a “cavity back,” increases the “moment
of inertia” of the club head, and so reduces club head rotation for off-center hits.
This was first established in a putter in 1966 by Karsten Manufacturing Company.
Named the “Anser” putter, it became the most successful putter ever both in terms
of sales and tournaments won. By the 1980s, the inventor, Karsten Solheim, had
introduced an even more significant innovation by using investment casting to cre-
ate cavity-back irons using the trade name “Ping.” According to Thomas (2011), the
second generation of cavity-back irons, the Ping Eye2, was used by Bob Tway in win-
ning the 1986 PGA Tournament. Now investment cast cavity-back irons are used
by almost all amateur players. The original forged “blade” irons, with the weight

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.
6 Introduction

concentration at the sole, are still used by a majority of professionals and highly
skilled amateur players. The reason typically advanced for this is that the blade irons
have better “feel” and provide better ball flight control. However, it seems much
more likely that these clubs are more suited for the downward angle of attack on the
ball and the resulting extra compression of the ball between the club face and the
inertial resistance of the ground.
Investment casting is just the modern term for the lost-wax process, with auto-
mation applied for low-cost manufacture. It allows parts to be cast with a very high
level of precision, including the very smallest features. This allowed Ping to cast
U-shaped grooves into their irons, which eventually brought the Ping develop-
ment in grooves to the attention of the USGA. As mentioned previously, golf balls
need to grip the face for high-spin generation, and they do this much better with
U grooves than the V grooves traditionally machined into forged irons. The main
issue became that if the U grooves are large enough with a very small profile radius
at the top edges, then the club face could grip the ball enough in deep rough to
still spin the ball well. This threatened to remove much of the penalty from errant
approach shots, which finally resulted in the USGA fighting a multimillion dol-
lar lawsuit against the Ping Corporation. One of the best lines in golf literature
is in Just Hit It (2008) by Frank Thomas in which he writes, “I came to think of
Karsten (Solheim, Founder of Ping) as a friend, even though he sued me personally
for $100 million.”
It was only a matter of time before investment casting would be used to cre-
ate hollow drivers with greatly increased moment of inertia. The first successful
introduction of this driver design, using investment cast stainless steel, was by the
Callaway Golf Company in the late 1980s. This led them to dominate the golf club
market for 10 years. By 1995 they had gone from $5 million to $500 million in sales.
At this point they introduced the Big Bertha, investment cast using high-strength
titanium alloy to produce even larger driver heads with consequently greater
moment of inertia. The high strength-to-weight ratio of the titanium alloy was
accompanied by the unparalleled lightweight spring quality of the material. This
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

produced additional performance improvements, which will be discussed follow-


ing a brief diversion.

A BR IEF DISCUSSION OF MOM ENT


OF INERTI A (MOI) IN GOLF
We will pause for a “moment of inertia” to describe what it is and why it is so impor-
tant in club head design. It is a consequence of Newton’s laws that the resistance
to twisting, or rotational inertia, of any item is proportional to its mass and to the
square of its distance from the center of the rotation. This is why MoI is expressed
in units of gram-centimeter squared. It is easiest to demonstrate this with a piece of
pipe being rotated about the axis through the center. If the tube has a thin wall, then
all of the mass can be considered to be at a distance equal to its radius from the rota-
tion axis. This sets us up for a simple experiment. We apply a torque to the tube and
measure its rotation speed. Next we get a piece of tube with double the radius and
half the wall thickness, which therefore weighs the same. We find that for the same

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.
7 Introduction

torque, it spins at only one quarter of the speed; that is, the radius doubled makes
the resistance to rotation increase by 2 squared or 4.
So the reason for making driver heads large is to move the 0.44 pound typical
head weight as far away from the center of the club head as possible. The torque pro-
duced by off-center ball strikes then rotates the head backward less and so projects
the ball forward more. The same applies to off-center strikes with irons and put-
ters where perimeter weighting has pushed the mass outward from the center, thus
increasing the MoI.

BACK TO THE HISTORY


The carbon fiber, reinforced, light-weight shaft was developed in 1969 by the
Shakespeare Sporting Goods Company, for whom Frank Thomas (2011) was then
Chief Designer. Since that time, carbon-fiber shafts have undergone continuous
development, particularly with respect to improved stiffness per weight; shafts are
now available in the 50 to 60 gram range. Despite claims to the contrary, this has
resulted in only the tiniest of increases in swing speed, and likely no increase in ball
speed, when compared to even the heavier hollow steel shafts, which are now almost
never used in drivers. However, the fiber layup construction of shafts has allowed
the stiffness distribution along the shaft to be customized for different player abili-
ties. In particular, regions of increased flexibility are positioned toward the end of
the shaft to promote forward rotation of the club head for increased “dynamic” loft.
These issues will be discussed fully in Chapter 3.
As mentioned previously, in 1995 the Callaway Company introduced the first
high-strength titanium alloy driver with a thin diaphragm-spring face. It is inter-
esting that Callaway adopted titanium alloy as a means of further increasing the
volume, and therefore the moment of inertia, of their hollow Big BerthaTM range of
drivers while keeping the weight within the accepted 0.44 pound range. Popular
lore has it that the increased ball speed that resulted from the excellent spring qual-
ity of the titanium alloy, and allowed the thin driver face to act like a trampoline,
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

was an unexpected bonus. This may in fact be true, because Callaway did not file a
patent on the trampoline effect, with disastrous consequences for the company. By
1997, Callaway, after spectacular growth from a few million dollars to almost one
billion dollars in sales, was selling more of all golf products than almost all other
golf companies combined. However, by 1998, eighteen oversize titanium drivers,
from thirteen manufacturers, were competing for market share, and the Callaway
market share was in decline. It is interesting to conjecture that, with a robust patent,
Callaway would likely have dominated the entire golf market up to this time.
As Thomas (2011) reports, one of Callaway’s main competitors informed the
USGA that they were finding unusual performance results in their testing of the
titanium club. This performance improvement was the subject of a 2001 article
by Michal and Novak, the main focus of which was to demonstrate that even bet-
ter performance might be possible through the use of amorphous metals, often
called liquid metals, in driver head design. Michal and Novak used a measure of
trampoline, or more correctly diaphragm, spring quality (Dieter 1983, Ashby
2005, Dewhurst and Reynolds 1997) to compare different materials for club head
design. Michal and Novak showed that liquid metals, particularly VitreloyTM by

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.
8 Introduction

Liquidmetal Technologies (2011), had the potential for superior golf club face per-
formance. However, attempts to produce a golf club with this material were unsuc-
cessful, apparently through a combination of fatigue-failure issues and difficulties of
bonding the faces to driver bodies. When the list of candidate materials is reduced
to those already applied successfully in club heads, titanium and beryllium copper
are shown to be equally best in class for driver diaphragm spring faces. However,
when the desire is to combine large head size with diaphragm spring quality, mate-
rial density must also be considered, which puts high-strength titanium alloy alone
at the top of the performance list for driver head design (see Chapter 2 of Design
for Manufacture and Assembly, Boothroyd, Dewhurst, and Knight 2011). It is inter-
esting to note that later on, the Callaway Corporation, apparently forgetting the
reason for the pioneering breakthrough they had made, developed a carbon-fiber
reinforced driver head and face for which diaphragm spring quality is far inferior to
high-strength titanium, which was unsuccessful in the market place.
Much more recently, golfer performance has been considerably enhanced
through the invention by Fredrik Tuxen (TrackMan A/S, U.S. Patents 2007,
2008, 2009)  of a radar monitoring system of ball flight, and his founding of the
TrackmanTM Company in Denmark. Trackman performs precise measurements of
club head and ball velocities, launch angles, ball spin rates, and the tilt angles of the
ball spin axis, in addition to monitoring the complete ball trajectory. It has proved
to be a powerful training tool, allowing players to try different combinations of club
loft angle, shaft stiffness, and strike attack angle, along with variations of their swing
mechanics, to obtain optimum conditions for ball flight. This complemented the
development of new multilayer balls designed to allow high launch angles in driv-
ing without the generation of excessive backspin. Thomas (2008) envisioned these
golf ball developments and, in 1999, proposed an optimized overall distance rule.
The intention was to have a maximum overall distance for all balls but using the
particular optimum launch conditions for each manufacturer’s ball. This proposal
was never adopted, but a rule for overall distance under a standard set of ball flight
conditions was adopted in 2004 by both of the governing bodies of golf, the R&A
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

(Royal and Ancient Golf Club of St. Andrews) and the USGA, referred to with
slight abbreviation going forward as R&A/USGA.
Tuxen has published a number of articles (2007–2010) over recent years con-
taining comprehensive sets of data on ball launch conditions and ball flight, together
with a number of empirical rules relating fundamental aspects of ball striking to ball
flight. With the permission of Tuxen, these have provided invaluable data for this
work, in some cases providing validation of analytical predictions, for others allow-
ing estimations to be made of model input parameters. Reference will be made to
individual newsletters in later chapters.
For a given impact velocity with a given club head mass, the principal reasons
for loss of distance when driving a golf ball are the loss of energy in the golf ball
itself as well as the twisting of the club head resulting from off-center hits. A golf
ball is far from perfectly elastic. When deformed during impacts of approximately
100 miles per hour, with even the latest high-performance balls and titanium driv-
ers, 30 percent or more of the energy transferred to the ball in deformation is typi-
cally lost to internal friction in the ball polymer structure rather than recovered
in extra ball speed. The original developers of golf equipment, contending with

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.
9 Introduction

much less efficient balls than those currently used, presumably discovered by trial
and error that greater distance could be obtained by striking the ball with a softer
club, which by sharing the deformation of impact would necessarily reduce the
ball’s deformation and consequently the amount of energy loss. The best material
for this purpose, with sufficient strength to sustain the impact loads, was found to
be hardwood, shaped so that the face cuts across the growth rings. This ensures
that the stiffness in the direction of impact is the lowest possible; in fact, less than
8 percent of the stiffness at right angles to the wood fibers, as, for example, that for
a baseball bat (Wood Handbook 1999). Compression tests carried out by me, on
specimens cut normal to the face of persimmon wood blanks from the Louisville
Golf Company, gave a stiffness modulus in the direction of impact of only 90,000
pounds per square inch. This is considerably less than the modulus value of 130,000
pounds per square inch used by Michal and Novak (2001), on the basis of which
they predicted that the maximum force from a 100 miles per hour impact, with
a wooden driver, would be 3,300 pounds; even less than their predicted value of
3,500 pounds for a diaphragm face modern titanium driver. However, Michal and
Novak used a “static” rather than “dynamic” solution, so we should only accept the
relative magnitudes of the two values.
The problem with the performance of hardwood drivers is that significant energy
is lost in internal friction between the wood fibers as well as in friction inside the
ball; with the modern ball, this may be a zero-sum tradeoff. Moreover, as discussed
previously, the distribution of the approximately 0.44 pounds of mass throughout
the bulk of the club head gives a very low moment of inertia compared to modern
hollow titanium drivers. Following the recognition of the increased performance of
the hollow titanium Callaway driver heads in the late 90s, the R&A/USGA estab-
lished a new equipment rule that the coefficient of restitution resulting from a 109
miles per hour impact with a golf ball should not exceed 0.83. To enforce this rule,
both the clubs and golf balls had to be subjected to independent tests. Both tests
use ball cannons equipped with ballistic screens to measure the rebounding ball
velocity after impact. For the ball test, a standard titanium alloy circular test plate
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

was adopted. The plate is 4 inches in diameter, with a 3-inch diameter inner region
0.115 inches thick, to form a spring diaphragm; and with an outer thick flange to
provide a total weight of 0.44 pounds, equal to the typical driver head weight. The
plate was made from the same titanium alloy and designed to give the same impact
performance as the Callaway titanium drivers already in the market. Balls exceed-
ing 0.83 CofR when fired at the plate at 109 miles per hour are deemed noncon-
forming. With this test in place, balls with the maximum 0.83 CofR value could
then be fired at the sweet spot (the point directly in front of the center of mass) of
driver heads to check that the CofR did not exceed 0.83. In addition to the CofR
rule, the R&A/USGA now restrict the volume of drivers to 460 cubic centimeters
and the moment of inertia component about the vertical axis (vertical MoI) to 5,700
gram-centimeter squared. Almost all manufacturers now supply their drivers with
the maximum volume. However, attempts to increase the vertical MoI into the
5,000+ gram-centimeter squared range, with rectangular-shaped driver head pro-
files, did not meet with success in the market; and the manufacturers seem to have
now settled on vertical MoI values in the area of 4,600 gram-centimeter squared for
460 cubic centimeter clubs.

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.
10 Introduction

While much attention has been focused by the R&A/USGA on the impact
efficiency of clubs and balls, surprisingly nothing in the rules concerns the aero-
dynamic performance of golf balls, except the overall distance restriction. At the
time of the introduction of the CofR rule in 2002, the R&A/USGA overall dis-
tance standard (ODS) required that a ball struck by a conforming (0.83 CofR)
driver at 109 miles per hour should not exceed 297 yards in total distance, includ-
ing bounce and roll, under carefully controlled standard conditions. In 2004, to
reflect the increased driving distance of professional golfers, the test was extended
to the requirement that a total distance of 320 yards should not be exceeded with
an impact of 120 miles per hour. Currently, the average PGA Tour player drives the
ball for a carry distance of 269 yards, with an additional average bounce and roll
distance estimated in Chapter 3 to be 40 yards. This average overall distance of 309
yards is achieved at an average impact speed of 112 miles per hour, so it must be
assumed that the longest hitters are routinely driving further than the R&A/USGA
limit. The loop hole in the rules is that the ODS is for a defined launch angle and spin
rate of the ball. Therefore, manufacturers are free to design balls that may exceed the
ODS but with different optimum launch angles and spin rates than defined for the
ODS. If Thomas’s proposed Optimized ODS had been adopted, every new ball on
Tour would have been wind-tunnel tested to determine its optimum launch condi-
tions, which then would have been used to assess overall distance.
Chapter 2 is devoted to the aerodynamics of golf ball flight. However, it is fitting
to provide a short historical introduction to this work here.
From the substantial literature on golf ball flight, two papers are of particular
interest for the combination of careful experimentation and extensive databases
they provide. The first is an exhaustive experimental assessment of the perfor-
mance of golf balls, carried out in 1976 at Imperial College, London, by Bearman
and Harvey (B&H; 1976). They carried out a series of wind-tunnel tests of dimpled
balls, equipped with internal motors to provide varying spin rates. The authors mea-
sured lift and drag forces and produced tables of lift and drag data covering the range
of golf ball speeds and spin rates experienced in play. They tested two different 1976
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

Uniroyal brand balls: one with circular and one with hexagonal dimples. They dem-
onstrated close agreement between ball flight distances of the hexagonal-dimple
balls, as predicted using their wind-tunnel data, and measured sets of launch angle,
ball velocity, initial ball spin, and flight distance. Because of the distance validation,
this extensive data set provides a benchmark in time against which to compare the
aerodynamic performance of the modern golf ball. Two decades later, Smits and
Smith (S&S; 1994) used a higher-speed wind tunnel, mounted golf balls on slender
spindles, and measured lift and drag forces and the rate of spin decay, for a wide range
of air speed and spin rates applicable to the driver through short iron shots. Smits
and Smith obtained lift data that “in all respects are similar to the data obtained by
Bearman and Harvey,” although their values were higher than the hexagonally dim-
pled ball data of B&H by a constant increment of 0.04. S&S also obtained “broad
agreement” with the drag coefficient data of B&H, although their results indicated a
“stronger dependence on spin-rate.” This is to be expected; if golf ball designers had
managed to tweak the dimple spacing and profiles to achieve greater lift, then it is
likely this would have been accompanied by an added amount of induced drag. For
the present study, the B&H lift and drag data were incrementally adjusted by the

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.
11 Introduction

40
30
(yards)
20
Callaway robot tests
10 B&H Hexagonal ball data
S&S ball driving data
0
0 50 100 150 200 250
(yards)

Figure 1C  Trackman measurements of robot driving tests with predictions using the Bearman
and Harvey (1976) lift and drag coefficients and the Smits and Smith (1994) ball aerodynamic
drive model; data on the Great Big Bertha II courtesy of Callaway Corporation.

writer until best agreement was obtained over the range of PGA and LPGA Tour
player trajectories. For the drive, the S&S model was found to give better agreement
with current premium balls. The nature of the lift and drag forces acting on the ball
is explained in Chapter 2.
Figure 1C shows actual ball flight from robot testing, as monitored by the
Trackman radar system. The tests were carried out, as a demonstration for the
writer and one of his colleagues, by the Callaway Golf Corporation using a con-
forming golf ball and the Great Big Bertha II titanium driver. From an average of ten
100 miles per hour hits, the measured results of the drive were:  initial ball
speed = 154.3 miles per hour; launch angle = 12.9 degrees; initial backspin = 3,106
revolutions per minute; maximum height = 38.3 yards; and carry distance = 245.1
yards. The dashed line (in Figure 1C) shows the predicted trajectory obtained
by the writer, using the original Bearman and Harvey lift and drag data for the
hexagonal-dimple, 1976 Uniroyal ball. Calculations were carried out as described
by Bearman and Harvey (1976), but using a spin decay rate as later determined
experimentally by Smits and Smith (1994).
With very minor adjustments, the S&S aerodynamic model simulated the
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

Trackman trajectory almost perfectly as shown by the square symbols in Figure 1C.


Before concluding this chapter, it is difficult to overlook the quite amazing per-
formance of the 1977 Uniroyal hexagonal-dimple ball. With all of the advances
in the understanding of fluid flows, and the development of sophisticated com-
putational fluid dynamics software systems for advanced aerospace design over
the last two decades, it would have been natural to expect a little more. However,
this comparison does not tell us anything about the ball striking behavior of
the Uniroyal ball compared to the modern premium ball. It would be wrong to
assume from these flight comparisons that the Uniroyal ball would perform as
well in actual play, and there is ample evidence to show that it certainly would not.
Likewise, the robot testing, with ten perfectly centered face strikes, gives no indi-
cation of its performance in actual play, particularly with a high-handicap player
whose ball strikes would be scattered quite widely from the face center. Dealing
with issues such as these, and of course different issues arising for every aspect of
the game, is the goal of this work.
As mentioned at the beginning of this Introduction, Trackman data on PGA and
LPGA Tour players will be used as the science “examples” throughout the text. This

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.
Table 1A  Average Trackman test results and modeling for PGA tour players

Club Attack Ball speed Vertical Spin rate Max Landing Carry
speed angle (mph) launch (rpm) height angle (deg) (yards)
(mph) (deg) (deg) (yards)

Driver 112.0 −1.3 165 11.2 2,685 31 39 269


3-wood 107.0 −2.9 158 9.2 3,655 30 43 243
5-wood 103.0 −3.3 152 9.4 4,350 31 47 230
Hybrid 100.0 −3.3 146 10.2 4,437 29 47 225
3-iron 98.0 −3.1 142 10.4 4,630 27 46 212
4-iron 96.0 −3.4 137 11.0 4,836 28 48 203
5-iron 94.0 −3.7 132 12.1 5,361 31 49 194
6-iron 92.0 −4.1 127 14.1 6,231 30 50 183
7-iron 90.0 −4.3 120 16.3 7,097 32 50 172
8-iron 87.0 −4.5 115 18.1 7,998 31 50 160
9-iron 85.0 −4.7 109 20.4 8,647 30 51 148
PW 83.0 −5.0 102 24.2 9,304 29 52 136
PW = pitching wedge.
Data reproduced with permission from Trackman A/C, Vedbaek, Denmark.

Table 1B  Average Trackman test results and modeling for LPGA tour players

Club Attack Ball Vertical Spin rate Max Landing Carry


speed angle speed launch (rpm) height angle (deg) (yards)
(mph) (deg) (mph) (deg) (yards)

Driver 94.0 3.0 139 14.0 2,628 25 36 220


3-wood 90.0 −0.9 132 11.2 2,705 23 39 195
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

5-wood 88.0 −1.8 128 12.2 4,501 26 43 185


7-wood 85.0 −3.0 123 12.7 4,693 25 46 174
4-iron 80.0 −1.7 116 14.3 4,801 24 43 169
5-iron 79.0 −1.9 112 14.8 5,081 23 45 161
6-iron 78.0 −2.3 109 17.1 5,943 25 46 152
7-iron 76.0 −2.3 104 19.0 6,699 26 47 141
8-iron 74.0 −3.1 100 20.8 7,494 25 47 130
9-iron 72.0 −3.1 93 23.9 7,589 26 47 119
PW 70.0 −2.8 86 25.6 8,403 23 48 107
PW = pitching wedge.
Data reproduced with permission from Trackman A/C, Vedbaek, Denmark.

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.
13 Introduction

data is presented here in Tables 1A and 1B as an easy to find reference for recurring
discussions of Tour player performance.
Ball flight or “carry” distances shown in the tables are not as large as might have
been expected. This suggests that for a player to be “on Tour,” the peak level of
achievement in the game, the major requirement is not ultra-long hitting, so it can
only be consistency of ball striking, coupled with accuracy on the putting surface
and of course a high tolerance for stress. However, we should certainly not underes-
timate the importance of distance combined with accuracy off the tee, for separat-
ing the very good from the very best.
On to the ball that actually does fly by creating its own propulsion system.
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.
Copyright © 2015. Oxford University Press, Incorporated. All rights reserved.

Dewhurst, Peter. The Science of the Perfect Swing, Oxford University Press, Incorporated, 2015. ProQuest Ebook Central,
http://ebookcentral.proquest.com/lib/itcr-ebooks/detail.action?docID=4083501.
Created from itcr-ebooks on 2021-11-10 00:29:14.

You might also like