Sie sind auf Seite 1von 144

Introduction I am a physicist and a keen golfer.

A few years ago I was under doctors orders to stay away from golf for at least 3 months in order to recover from a shoulder injury. So instead of playing, I trawled through golf books and papers to figure out what actually happens in the golf swing. After many months, I had a good pile of paper describing a lot of things about golf, but nothing simple to tell me how the swing really works. Even the best of the classic books like Cochrane and Stobbs, Daish, and Jorgensen fail on this point. Once I was back playing, the problem continued to nag. The underlying problem with golf-swing physics is the complexity of the math. The equations explain everything and nothing. Sure it is possible to solve the equations, produce animations, and carry out numerical experiments, which is what most researchers have done, including me. However, the equations are so complex that it is not possible to look at them and say, Ah ha! I understand. Eventually I found the simplification that provides the insight, and this presentation is the result. The presentation builds the whole of the golf stroke including the swing and the club-ball collision. The key points of technique identified here are simple but so counterintuitive that they are hard to believe and execute when you are standing over a ball. The physics explanation helped me to defeat my intuition. Im in my 50s now and 30 years of bad wiring in the brain is hard to change, but over a period of two years I have gained 30-50 m distance on my drives, and improved the accuracy on all of my shots, all with less effort. Hopefully these pages will help you to do the same. A careful reading of the presentation will take about an hour, and for most people its probably necessary to view the animations to really understand how the swing works. If you dont have an hour, you can try reading the technique section by itself, but you may find some of the advice hard to believe.

The golf swing is a combination of two physical processes: most obviously, a collision between the club and ball, and perhaps less obviously, the unfolding of a simple machine called a double-pendulum. Although simple, the double pendulum has very complicated behavior. In recent years, it has received a lot of attention because is a simple example of a system exhibiting chaotic behavior (which may explain some of my golf shots). The double pendulum also turns up in a lot of different places because it is an efficient way of transferring energy from a heavy slow-moving object to a light fast-moving object:

Baseball, tennis, hockey, and wood chopping are obvious examples. Football (kicking), javelin, and discus are not so obvious examples. Old war machines: Trebuchet (like a catapult), sling, atlatl or spear thrower (also known as the woomera in Australia). Builders and shipping cranes (especially on ships where it causes trouble), and the threshing flail.

Outline of the Article 1. The collision o 2-ball collision, as a model of the club-ball collision o 3-ball collision, a simple analogy of combined collision and double pendulum 2. The swing o The double pendulum o Combined collision and double pendulum 3. The good and bad of technique o A driven double pendulum o How is energy transferred? o The good the effect of wrist cock (lag) o The bad the effect of wrist torque (using the hands) o Video examples of good golf swings o Video examples of trebuchet swings 4. Golf Technology o Club-head mass o Shaft length o Coefficient of restitution o Shaft mass

References: 1. Theodore Jorgensen, The Physics of Golf (Springer Verlag, New York, 2nd ed, 1994) 2. C. B. Daish, The Physics of Ball Games Parts I and II (English Universities Press, London, 1972) 3. Alistair Cochran and John Stobbs, Search for the Perfect Swing (Triumph Books, Chicago, 1968)

Golf Swing Physics 1. The Collision Guest article by Rod White -- December 2008 This section is in two parts. The first subsection, on the two-ball collision, provides a simplified explanation of what happens in the collision between the clubhead and ball. The second subsection, on the three-ball collision, considers what happens when two two-ball collisions are combined. This may seem like a bit of a detour, but hang in there; it leads to an explanation of why the club head is so much heavier than the ball. More importantly, it captures the essential features of the compromise required to maximize the overall energy transfer in the golf stroke. Two-Ball Collision Let us look at a moving ball colliding with another ball that is initially stationary; our aim is to get as much kinetic energy as possible into the second ball. In an efficient collision the first ball stops. This is the criterion for 100% efficiency. Any movement of the first ball after the collision represents kinetic energy that has not been transferred to the second ball. Case 1: small ball collides with large ball

This slide shows an animation of the collision. After the collision, the large ball moves slowly, and the little ball has recoiled and is moving slowly backwards. As evident from the movement of the small ball after the collision, not all of the kinetic energy of the small ball has been transferred to the large ball. The collision is not 100% efficient. Additional note: In these animations it is assumed that the collisions are ideal, there is no energy lost through compression of the balls (i.e., the

coefficient of restitution = 1.0) Math note To a physicist, this page illustrates perfectly the laws of conservation of energy and momentum. Let's calculate the energy and momentum for each ball, before and after the collision. Before the collision: Kinetic energy: Momentum: T = M1V12 L = M1V1

After the Collision: Kinetic energy: Momentum: T = M1U12 + M2U22 L = M1U1 + M2U2

Where V is the velocity before collision, U is the velocity after the collision, and M is the mass of the balls. The conservation laws say the the "before" should be equal to the "after", for both kinetic energy and momentum. So we now have two equations, Tbefore=Tafter and Lbefore=Lafter. Since we know everything except the aftervelocities U1 and U2 , we should be able to solve the two equations for the two unknowns -- and we will below. Case 2: large ball collides with small ball

This time the animation has the larger ball colliding with the small ball. Again the collision is not 100% efficient, and afterwards, the large ball continues to move forward: it retains some of the kinetic energy. Note that the little ball moves faster after the collision than the big ball did before the collision. We will come back to this point in a few slides.

Math note Above we saw how conservation of energy and momentum give us two equations in two unknowns. So it is now just algebra to calculate the velocities after the collision in terms of the initial velocity of ball 1 and the masses of the two balls. If we solve the equations, we get... Velocity of first ball after the collision is U1 = V1 M1 - M2

M1 + M2 Velocity of second ball after the collision is U2 = V1 2 M1 2 = V1

M1 + M2 1 + M2 / M1 Note: You have probably seen the equation for ball speed after collision with the clubhead, cited by many references, as: 1 + COR 1 + Mball / Mhead Let's look at the two-ball collision as a golf shot, with ball 1 as the clubhead and ball 2 as the golf ball. Our model so far has no energy loss, so the CoR is 1.0. Given this value of CoR and assignment of masses, the familiar equation for ball speed is exactly the same as the equation for U2, the speed after collision of ball 2. Now we know where the equation for ball speed comes from. Case 3: two equal sized balls in collision Vball = Vhead

This time the two balls have the same mass, and the colliding ball comes to a dead stop with the collision. The collision is now 100% efficient.

We can also make a couple of other observations about the collision -- and you can even see where they came from if you followed the math notes:

The condition for the first ball to stop is M1=M2, and the second ball takes on the original velocity V1. When the first ball is much more massive than the second, the velocity of the second ball approaches 2V1.

This graph plots the collision efficiency versus the ratio of the two masses. The efficiency is the fraction of the total kinetic energy transferred to the second ball. The graph shows the collision follows the Goldilocks principle: the colliding ball can be too big, too small, or just right. Just right is when the two masses are equal. So why does the typical golf driver have a mass of 200 g and the ball a mass of 46 g, surely the collision would be more efficient if the two masses were the same? We will answer that question shortly. Math note The equation of this curve is Efficiency = 4 M1 M 2

(M1 + M2)2 Three-Ball Collision Now, lets look in detail at the three-ball collision. It might look like a bit of a detour, but we will see shortly that two successive 2-ball collisions illustrates an important compromise that occurs in the golf swing.

Remember, when we ran the animation of the large ball colliding with the small ball, how the small ball moved faster than the large ball? Although its not easy to tell, the small ball now moves faster than before. The intermediate ball makes the transfer of energy from the large ball to the small ball more efficient.

This graph plots the efficiency for the two separate 2-ball collision that occur within the 3-ball collision; first M1 with M2, and then M2 with M3. In the graph I have assumed that the big ball has a mass of 1 kg, and the small ball has a mass of 50 g. The mass of the intermediate ball can take any value as indicated on the horizontal axis of the graph.

The right hand dotted curve shows the efficiency of the collision between M1 and M2, with the mass of M1 equal to 1 kg. As expected from our observations with the 2-ball collisions, the energy transfer is most efficient when the intermediate mass is also 1 kg. The left hand dotted curve shows the efficiency of the collision between M2 and M3, with the mass of M3 equal to 50 g. It is most efficient when the intermediate mass is 50 g. After the first collision, the intermediate mass carries a given percentage of the energy, and then this fraction of the energy is reduced by a further percentage in the second collision. The efficiency of the overall three-ball collision is therefore found by multiplying the efficiency the two collisions, as given by the two dotted curves. The result is the solid red curve. Clearly, the best overall efficiency is obtained when the intermediate mass is in between the other two. If we read the mass off the graph we can see it is a little greater than 200 g. Math note The velocity of the third ball is given by: U3 = 4 M1 M3 V1

(M1 + M2)(M2 + M3) Differentiation with respect to M2, and setting the result to zero gives the condition for the maximum energy transfer:

In other words, the optimum intermediate mass is the geometric mean of the other two. To improve the efficiency even further we could introduce a fourth ball, or an infinite sequence of balls, with the masses graded between the large and small masses. Now the efficiency approaches 100%. Practical examples of this type of exponentially graded coupling include the bullwhip, horn loudspeakers, and graded electromagnetic absorbers used in stealth technology. Golf Swing Physics 2. The swing

Guest article by Rod White -- December 2008 Double-Pendulum Swing Now we move on the model of the swing itself. Conventionally, physicists model it as a double pendulum: one pendulum tacked onto the end of another.

In this animation we show a double pendulum composed of a large blue mass representing the arms of the golfer, and a smaller red mass representing the clubhead. There is also a thin string connecting the clubhead to the hub. Initially the system rotates at a constant speed, and the string stops the clubhead from swinging out. No energy is put into the system for this model; the system is rotating to begin with, and continues to rotate under its own inertia. Since it is held in the same shape by the string, the moment of inertia stays the same so the speed stays the same. An essential feature of the animations is that no external forces or torques are applied to the system, and the connection between the two arms of the pendulum is simply a hinge. Now watch what happens when the string is removed. As the red mass swings out, the blue mass slows down, and the red mass accelerates. When the system is fully extended (near the point where the head would hit the golf ball) the blue mass comes to a dead stop. For the physicist, this is the magic moment! The blue mass coming to a dead

stop means that all of the kinetic energy that was in the blue mass has been transferred to the red mass (the club head); the swinging process is 100% efficient. Most importantly, there were no external forces applied to make this happen the system rearranged itself under its own inertia. If this was a golf swing, the golfer would not have to make any effort to make it happen. In this example, I chose the value of the red mass so the blue ball would come to a dead stop. In general this does not happen as we will see with the next few animations.

The second animation shows the double pendulum again, but with a heavy clubhead mass -- heavier than before and therefore heavier than optimum. This time, the blue mass slows down more quickly than before, and when the system is fully extended the blue mass actually goes backwards (this tells us that some of the kinetic energy remains in the blue mass). This animation is similar to the two-ball collision when a small mass collides with the big mass and recoils.

The third animation shows the double pendulum again, but with a light clubhead mass. The distribution of mass is closer to a real golf swing than in the previous examples. This time, the blue mass slows down but not as much as before, and when the system is fully extended, the blue mass is still moving forwards. Again some kinetic energy remains in the blue mass. This animation is similar to the two-ball collision when a large mass collides with the little mass and keeps moving. Now although the motion of the double pendulum is complicated, we can simplify the problem by considering the initial motion when the string is in place, and later when it is fully extended. If we think about these two snapshots, and ignore the complicated movements as the red ball swings out, we can see a strong similarity between the two-ball collision and the golf swing as it unfolds. Once again we see the Goldilocks principle at work. The red mass representing the club head can be too big, too small or just right. We wont show you the curve just yet, but it is very similar to the efficiency curve for the 2-ball collision. The double pendulum is much more complicated than the simple collision, so the condition for a 100% efficient swing is much more complicated that the condition for a 100% efficient collision (which was M1 = M2). We give the equation for the optimum clubhead mass in the math note below. In addition to the mass of the blue ball representing the arms (M1), the condition depends on the length of the two arms of the pendulum representing the golfer's arm and the shaft of the club (L1 and L2), and the angle between the arm and the shaft (, which is the wrist cock angle).

The equation also tells us how to reduce the optimum clubhead mass and increase the efficiency of the swing:

Lengthen the shaft, and/or Increase the wrist-cock angle.

In fact the whole golf stroke: the unfolding of the swing, and the club-ball collision looks a lot like a three-ball collision. The equations of motion for the system are still complicated, but if we restrict ourselves to thinking only about the energy and momentum at these two points (the point the string is released, and the point of full extension), the system can be understood relatively easily. Remember that the energy and momentum are the same throughout the swing and through impact. The momentum must be the same; that's physics. The energy is the same because we have not yet considered energy losses in the collision (that is, the COR is 1.00 for our model). Math note

Here is the basis of the mathematical model I am using. The diagram shows how Ive defined the various angles, shaft lengths, etc. We are only considering the instantaneous motion of the system at the time of two snapshots:

a. The moment just before the string is released. At this moment, everything is rotating; nothing is flying out from the centre of rotation -- though it will the instant after the string is removed. b. The moment of impact, calculated to correspond to "complete release" -- a straight line from the golfer's center of rotation down the arms and the club to the ball. By picking these two snapshots, and having the system move under its own inertia, the system is very much simpler to analyze. Mainly, there are no radial terms to consider; only tangential terms involving angular velocities. Only two things you really need to note here, both of them at the moment the string is released:

The downswing angle , which measures how far the arms move, and The wrist cock angle , which measures the angle between the arms and the shaft of the club.

These are the simplified equations for the swing. (We still have not included the collision with the ball at this point.) They might still look horrific, but they are much simpler than the full equations, and they can be solved algebraically to give a simple expression for the condition for 100% efficiency. To read the equations you need to know that terms of the form mass x length2 measure a quantity called the moment of inertia. Moment of inertia measures the resistance to rotational forces (torques) in the same way that mass is a measure of resistance to a translational force. Everybody knows "F=ma". Well the rotational analogue is "torque = moment of inertia x angular acceleration". The Greek symbols with dots over them are angular velocities. Kinetic form: energies have the moment of velocity)2 inertia x (angular

Momenta have the form:

moment of inertia x angular velocity

The subscripts on the angles, i and f, indicate initial and final velocity. First, we require the kinetic energy and angular momentum to be the same for the two snapshots of the swing.

Note that the angular velocity of the two masses, beta-dot and alpha-dot, is the same in the first snapshot. As we did with the equivalent equations for the collision, we can solve these two equations to calculate the final angular velocities for the two masses. This is how the swing efficiency curve was calculated. However, it is more interesting to find out the condition for a 100% efficient swing. Ideally, we want the blue mass to come to a dead stop at impact, and this leads to a further simplification, as given by the second set of equations.

This pair of equations is much simpler and has one non-trivial solution: m1L12 + m2R2 = m2 (L1 + L2)2 The bottom line of this analysis is that the swing is 100% efficient when the moment of inertia of the system in the first snap shot equals the moment of inertia of the club in the second snap shot. In words, this last equation reads, the moment of inertia of the initial system (with the string in place) equals the moment of inertia of the club when fully extended. This is analogous to the condition that the M1 = M2 in the twoball collision. Most importantly, this observation tells us that the energy transfer occurs because of the changing moment of inertia of the club as it swing out. We can now calculate the optimum clubhead mass for this passive model by substituting the value for R, which is given by the cosine law (from diagram a above): R2 = L12 + L22 - 2L1L2 cos Which then gives us the required condition for optimum clubhead mass: m2 = m1 L1

2[1 + cos()] L2 Ice skater analogy of the golf swing The most important picture to keep in mind is the transfer of energy with unfolding. Think of this as the reverse of the ice skater effect. The ice skater initially has the arms extended and works to move them as fast as possible. The skater then pulls the arms in close to the body. This causes the body speed up so that the skater spins fast. It is important to note that the hands actually slow down in this process, but we only watch the body. The body is spinning faster, true. But, because the hands' radius of rotation is shorter than before, the velocity of the hands is lower than before. In the golf swing, the opposite occurs. The golfer initially holds the club close to the body using wrist cock, and works very hard to build up the kinetic energy in the body and arms. The golfer then allows the club to swing away from the body, so the body and arms slow down and the club speeds up. I emphasise the term allows, because it can be an entirely passive process; the golfer does not have to make the club swing out it happens naturally. In summary...

The ice skater: o First builds speed in arms, o then folds the arms close to the body, o to speed up the body The golfer: o First builds speed in body, o then unfolds the club from the body, o to speed up the club.

Model golf stroke = pendulum + collision We are now in a position to understand the basic principles underlying the golf stroke. There are two parts in the golf stroke: the swing and the collision. This graph is idealised we discuss the non-ideal bits shortly.

Part 2: the collision between the club and ball is exactly as we described for the two-ball collision. The efficiency is represented by the left hand of the two dotted curves, and the maximum efficiency occurs when the two masses are equal (golf ball mass is close to 46 g). Part 1: the swing is where energy is transferred from the golfer's arms and torso to the clubhead. This is represented by the right hand dotted curve. The right hand dotted curve for the swing efficiency is not exactly the same shape as that for the collision but very close. The exact position of the curve depends on a variety of factors that we havent considered in detail yet arm mass, arm length, shaft length, wrist-cock angle, and coefficient of restitution. The maximum efficiency in this case for the swing alone requires the club head mass to be about 1 kg (according to the complex equation given in the math notes). As with the 3-ball collision, the overall efficiency is given by the product of the two separate efficiency curves. We can now see why the clubhead is so much heavier than the ball. To maximise the efficiency of the total golf stroke, it must be midway between the optimum for the swing (1 kg) and the optimum for the collision (46 g). The right hand dotted curve is not exactly the same shape as the corresponding curve in the three-ball collision but very similar. Anything we can do to move the swing curve further left improves the efficiency of the golf stroke. The only two factors included in this model that are within the golfers control are:

The wrist-cock angle (make smaller). That is why many of the longest

hitters have a very acute lag angle well into the downswing. (We shall see this in the videos later.) Shaft length (make longer). But shaft length is limited by the rules, and in many cases further limited by the individual golfer's ability to control a long club. Long drivers routinely use long shafts, but they too are limited by their competition rules.

The model we have developed so far suggests that the optimum club head mass is about 150 g. In practice, real club-head masses are a little heavier. This is because, unlike our model (which was entirely passive), in a real golf swing, the golfer is constantly working to build up energy and momentum. Work done early in the swing is transferred efficiently, just as our model suggests. Work done late in the swing, when the wrists are un-cocked, is transferred less efficiently. The overall efficiency is therefore less than our model predicts in effect the average wrist cock angle is greater than we have assumed. There are other factors too:

Energy lost due to the compression of the ball as described by the coefficient of restitution means that the swing is less efficient, and the clubhead mass must be heavier to give the maximum ball speed. Our muscles also produce greater torques when they are moving slowly and this favours a slightly heavier mass. Finally,the optimum mass changes with the value of m1 that represents the mass of the golfers arms, and our value of 1 kg on a 0.66 m pendulum arm may not be representative of the moment of inertia of a typical golfer.

Golf Swing Physics 3. Technique Guest article by Rod White -- December 2008 In the previous sections we explained the principle of the golf swing, how the unfolding of the club from the cocked position causes rotational energy to be transferred from the arms and the body to the club, and that this unfolding can be passive requiring no effort from the golfer. Now that we understand what is happening, lets look more closely at a more realistic model of the golf swing, in the interest of clarifying technique. In the actual golf swing, the golfer is applying torque, throughout the swing, to the inner arm of the double pendulum -- by using

muscles in the torso to turn the shoulders. The improved model is a similarly-driven double pendulum with a few extra features that allow us to investigate aspects of technique. It will give us an opportunity to see what sort of things the golfer can do with the swing to improve his distance -- or screw things up altogether.

A Driven Double Pendulum The improved model can vary:


Torque applied to the inner arm of the pendulum, to model the work done by the golfer via the torso and shoulders Torque applied to the outer arm of the pendulum, to model work done by the golfers hands Wrist cock angle Release timing, to model the golfer releasing the club early or late during the first phase of the downswing Arm mass Arm length Club shaft length Club head mass Coefficient of restitution for the club-ball collision

For the moment we will focus on the aspects of technique that have the largest effect on the effectiveness of the golf swing we will look at the technological factors later.

The animation shows the swing of a moderately good amateur golfer with a sound golf swing. During the first part of the downswing, the golfer holds the club in a cocked position and accelerates the shoulders and torso. Initially some positive wrist torque is required to stop the club from being pulled into the golfers neck (the hub). Remember the passive, steadily rotating model on the previous page? There, a string (providing negative torque) was needed to keep the club from swinging outward. Here, during the initial build up of speed, some sort of "brace" (providing positive torque) is needed prevent the club from being pulled inward. The positive torque required to brace the club falls rapidly as the club accelerates. When the positive bracing torque falls to zero, the club can be allowed to swing out -- ending the first phase. The second phase of the downswing occurs as the club swings out. If the golfer lets the club swing out when the bracing torque falls to zero, then this is described as a swing with a natural release. If the golfer holds the club in the cocked position for a short while longer, this is described as a late release. If the golfer releases the club early, the club will swing in towards the neck for a small moment and then swing out. We wont look at the effect of release timing because to a good approximation release timing has no effect. During the second phase the golfer continues to turn his body and arms, but no torque is applied via the hands they are no more than a hinge during this phase. This model will be the starting point for all our future calculations. The full numerical model includes a number of variable factors, as indicated in the list above. The animation shows the solution for the swing we have just described. Important note: By pure coincidence (perhaps), the golf swing can be executed with the natural release. This is not necessarily true for all stick-and-ball sports, not even if the stick is a golf club. Consider:

With baseball swings, the natural swing time is much shorter because

the bat is shorter, and the base-baller must restrain the bat using negative wrist torque to stop it from swinging out early. With professional long-drive golfers, the shaft length is longer (up to 50 inches), the natural swing time is much longer, and a swing with a natural release is anatomically impossible. Professional long drivers must use positive wrist torque (forcing the club out) to complete the swing. The normal golf swing does not require positive or negative wrist torque during the second phase of the downswing; therefore the hands are passive, and the golf stroke can be more accurate with fewer muscles involved.

How is the energy transferred?

We have discussed the golf swing in terms of the conservation of energy and momentum, and showed that the energy is transferred to the club as the swing unfolds, but what actually happens where are the forces that make this happen?

The figure shows a stroboscopic view of the golf swing. Have a close look at

the direction of the clubhead midway through the swing this is indicated approximately by the red arrow. Now look where the hands move at the same time the blue arrow: in a different direction! Obviously the hands and clubhead cannot continue to move in different directions, they are restrained by the fixed length of the shaft. The diverging directions of the club and hands results in a large tension in the shaft. The tension pulls against the club head causing it to accelerate, and pulls against the hands causing them to decelerate. It is the differing directions of the hands and club that are ultimately responsible for the energy transfer. In a professional golfer's swing, the tension peaks above 500 N (50 kg equivalent, or over 100 pounds). During this phase of the swing the rate at which energy is transferred to the club peaks at about 5 kW (or almost 7 horsepower). Now well take a look at the factors that affect the effectiveness of the swing. The two big factors are the wrist cock angle and the wrist torque. Since greater wrist cock increases the divergence of the trajectories of the hands and the clubhead, we can expect greater wrist cock (smaller wristcock angle) to improve the swing. It is less obvious whether wrist torque -usually described in golf as "hand action" -- helps or hurts. Let's look in more detail at the effects of wrist cock first. Wrist Cock Angle

The figure to the left shows the clubhead speed versus downswing angle (the angle between the arms and the body) for three different wrist cock angles. As expected, increasing the amount of wrist-cock (reduced angle between the arms and shaft) increases the efficiency of the swing. The key point is that the peak speeds all occur at a very similar downswing angle, showing that the swing timing is almost unchanged. The golfer expends the same effort for all three swings, yet we see a 10% increase in head speed resulting in a 10% increase in distance say 20 m for a 200 m drive -- with no extra effort.

The chart at the right plots the driving distance versus wrist cock angle, assuming all other aspects of the swing remain the same and that the ball is hit at the peak head velocity. The increase in distance that occurs with a decrease in wrist-cock angle between 110 and 70 degrees is about 20 m say 5m for each 10 degrees wrist cock. Note again that the wrist-cock angle is measured between the arms and the club, so a smaller angle corresponds to greater wrist cock, or greater lag as it is often called. Note again the distance is gained with no extra effort from the golfer the difference is purely one of technique. For those who are interested, I've estimated the distance from the clubhead speed using a formula from Cochrane and Stobbs. D = 3.75 x ball speed 25m where the ball speed is in meters per second. Anecdotal confirmation of this point from DaveT: In December 2009, I was playing in a foursome about my own age. We were all within a year or two of 70, and in relatively good shape for our age. Two of us had roughly a 90 wrist cock at the start of the downswing; the other two had almost no wrist cock at all. Throughout the round, it was telling that the two with wrist cock were roughly equal length; so were the two without wrist cock, but typically

30-50 meters back. Wrist Torque

Now we look at the effect of positive wrist torque during the second phase of the swing. The use of the hands is a very frequent flaw with amateur golfers. It is not uncommon to see the hands spread far apart (like a baseball grip), or the right hand adjusted so the thumb is behind the shaft through impact, or the right forefinger is set down the shaft. All this is done in the hope of pushing the head faster through the impact zone. In fact it has the opposite effect. In this figure, the wrist torque is expressed as a percentage of the shoulder torque. For the model Ive chosen, 10% corresponds to 1 kg.m of wrist torque in the model. This is a very large torque, but probably typical for male beginners who have yet to learn to let the club swing by itself. The graph shows that positive wrist torque causes the club to unfold early, and therefore causes the clubhead speed to peak early, and with a lower velocity. Common symptoms include a pronounced swishing sound that peaks before impact, drop-kicked shots (club ricochets off the ground before impact), shots with a high trajectory, and often problems with big high fades or slices. Researchers who have tracked the swing speed for golfers with a

range of handicaps find that only golfers with low single-figure handicaps or better come close to hitting the ball at the peak clubhead speed. For most golfers, the club is decelerating through impact.

The chart to the right plots the approximate driving distance versus wrist torque with almost all other parameters kept the same. Remember that wrist torque has two effects on clubhead speed. It (a) peaks at a lower clubhead speed and (b) peaks earlier in the downswing.

The blue curve assumes that the golfer changes his swing so impact still occurs at the peak. We shorten or lengthen the swing so that impact will occur at maximum clubhead speed. This golfer is then only bitten by (a) above. The red curve assumes that the golfer simply makes the same length swing no matter what the wrist torque. This golfer is then bitten by both (a) and (b). Negative wrist torque also costs distance because the clubhead speed peaks after impact (i.e., impact is at the black line in the curve above).

Even if we assume that the ball is hit at the peak head velocity (blue curve), the difference between a beginners swing (10% wrist torque) and a swing with no wrist torque is about 20 m in distance. More typically the beginner will take the same backswing as a low handicap golfer and lose the distance indicated by the red curve nearly 40 m! This is a very tough lesson, yet all of us have experienced the occasion when

we relax, try not to hit a ball too hard, and hit the best drives of our lives. Learn to relax, to shorten your grip, and not to use your hands. Many people have trouble believing that you do not need to use wrist torque to have an effective golf swing. But to prove a point, some stunt golfers use drivers with a section of rubber tube or dog chain replacing part of the shaft. They still hit the golf ball a long way -- in fact, much the same distance as with a proper shaft. With such a flexible shaft, there is no way that wrist torque can have any effect. Another good example is the trebuchet shown in a couple of the videos on the next page, a medieval siege machine used to fling rocks into or over castle walls. From the physics point of view, the trebuchet is an upside-down golfer; the raised weight represents the torque applied through the shoulders, the long wooden beam represents the golfer's arms, and the rope sling represents the shaft of the golf club. Remember that almost all the energy transfer to the club is due to tension in the shaft; the shaft does not need to be stiff, because the vast majority of the force it transmits is along the length of the shaft. If the club had a perfectly flexible shaft -- like the rope sling of a trebuchet -- then there is no way to apply wrist torque to get any action from the clubhead. Yet the trebuchet was a very effective siege weapon. It was the best, most powerful catapult in warfare for centuries until it was replaced by gunpowder-powered cannons, demonstrating that "wrist torque" isn't all that important. See the videos on the next page for visual demonstrations that a trebuchet can sling things a long way with no "wrist torque" at all.

Summary for Technique Work done by the golfer builds up kinetic energy in the torso, shoulders, and arms. This is then transferred via tension in the shaft as the club and arms unfold away from the golfers body.

The good: the greater the fold (wrist cock) the more efficient the transfer of energy from the body to the club.

The bad: the greater the wrist torque (use of the hands) the earlier the club unfolds and the less energy is transferred to the club.

These two effects, the negative effect of wrist torque and the positive effect of wrist cock, account for most of the 70 m difference between the beginner and the scratch golfer. These effects are also counterintuitive not what the beginner golfer expects. This perhaps explains why a good golf swing is so hard to learn. Another factor making a good swing hard to learn is that it is mentally difficult to hold onto the club firmly while not holding the wrists firmly. It is curious that most people, when asked to throw a golf club as far as possible, would swing the club around their shoulders without using wrist torque, and this is exactly the action required for a good swing. Swinging a club loosely around your shoulders as if you were about to throw it will help to train your brain to not use your hands. I have also found it helpful to visualise throwing the club through the impact zone. In fact a full vigorous swing around your shoulders like a baseball swing, including hip and shoulder movement, captures all of the important parts of the swing. One of the benefits of the overlap grip is that it keeps the combined length of the hands short and the right hand weak (for a right-handed golfer). This enables the golfers to grip the club firmly, but limits the ability to apply wrist torque. Math note I have not given the full equations for the numerical model here. As I indicated in the introduction, they are too complicated to yield any insights directly at least not for me. Also, they have to be solved numerically because there is no analytic solution. If you want to experiment with the equations, the full version can be found in Appendix 4 of Jorgensens book. For the analysis here I neglected gravity, assumed constant shoulder torque, and assumed constant wrist torque during the second phase of the downswing. This allows the equations to be partially integrated analytically and reduces the number of dynamic variables in the numerical integration from 4 to 2. For further information see the paper by Pickering and Vickers, or the supplementary EPAPS document associated with my paper it can be

found quickly if you Google EPAPS, golf swing. To do the integrations you will need a moderately good numerical integration algorithm. Applications like Mathcad, Maple, and Mathematica have very good integration routines. References: Theodore Jorgensen, "The Physics of Golf 2nd Ed", Springer Verlag, New York, 1994) W. M. Pickering and G. T. Vickers, On The Double Pendulum Model Of The Golf Swing, Sport. Eng., 2, 161-172 (1999) D R White, "On the efficiency of the golf swing", Am. J. Phys. 74, pp. 10881094 (2006) Golf Swing Physics 3+. Video Examples Guest article by Rod White -- December 2008 In this section well look at slow motion videos of some professional golf swings. The real golf swing is more like a triple pendulum with the beam between the neck and the shoulder forming the third arm of the pendulum. While watching the videos, remember the essence of a good golf swing is the reverse of the ice skater effect the transfer of energy with unfolding. The total power produced by the golfer (about 2.5 kW for a professional) comes from the biggest muscles in the body the upper legs and torso. The problem is how to transfer this power, from the muscles, through the shoulders, arms, club, and eventually, to the ball. The backswing

consists of a large shoulder rotation against the hips to pretension the muscles of the torso. minimises the moment of inertia of the arms and the club by folding the left arm against the shoulders and cocking the wrists. In all cases the down swing begins with the left arm folded almost parallel with the shoulders, and the club typically cocked to 90 degrees.

The downswing movement

leads with the large muscles of the legs and torso rotating and building momentum in the hips, with the whole folded system -- the upper body, shoulders, arms and club - moving almost as one. Once the rotation is firmly established, the arms swing out transferring energy from the body to the arms, then the club swings out, transferring energy to the club.

Tiger Woods Tiger's swing is the swing closest to the textbook swing of the golfers here. Things to note:

The long extended backswing to create a large shoulder turn and tension in the muscles of the torso. The system stays "folded" through the early part of the downswing. Excellent retention of wrist angle well into the downswing. Hands slow noticeably as he approaches impact. This is not "deceleration" in the golf instructor's [negative] sense; it is caused by the tension in the shaft exceeding the shoulder torque, and is evidence of the transfer of energy from the arms to the clubhead. Hand, wrist, and forearm strength are not a factor here -- except to be able to stay relaxed with the large shaft tension trying to yank the club from his hands. And the same comment applies to the other videos as well.

Sergio Garcia Sergio is not as tall as many pro golfers, so to hit the ball as far as he does he must have a more efficient swing. The distinguishing feature here is the large wristcock midway through the down swing. Things to note:

Even more shoulder turn than Tiger. As he starts the downswing, he actually increases his wrist cock angle. This "early release" of the club means that he's not using sufficient positive wrist torque to maintain even a 90 angle. Midway through the swing, the wrist cock is probably well under 60, so that work done during this part of the swing is transferred more efficiently to the club when the wrists unfold.

That is where his exceptional clubhead speed comes from.

Freddy Couples Fred has one of the most admired swings in the game, superficially an easy swing, but hes actually doing a lot of work very early. The important thing to note here is how Fred starts his downswing. At the top, his whole body is coiled, including his hips. His first move is a big rotation of the hips. Everything above the hips turns, but only because the hips are turning. Everything above the hips -- torso, shoulders, arms, and club -- are almost frozen, and just move with the hips. That is what we mean by "staying folded". The rest is pure swing -no hands. Of course Fred, like Tiger and Sergio, keeps his wrists cocked until very late in the downswing. Jamie Sadlowski Jamie is not a household name for most golf fans -- unless they are fans of long driving. Jamie, at only 165 pounds, has been the world champion of this sport for the past two years (2008 and 2009) with record drives exceeding 400 yards. How does he get the clubhead speed to accomplish this against much bigger, more-muscled competitors? Jamie is an incredible athlete with extreme flexibility, and none of us could hope to reproduce this swing. But because the downswing is so long, it shows very clearly the ideal of staying folded. As a long driver, all of his actions exceed those of PGA tour golfers: an incredible shoulder turn and even more incredible wrist cock. He also has the advantage of a long shaft on the driver (see next section on technology). Many people have trouble believing that you do not need to use wrist torque to have an effective golf swing. But there are some good examples from outside golf illustrating this point. Perhaps the best example is the trebuchet, a medieval siege machine used to fling rocks a long way. YouTube has many videos of trebuchets. Let's see what we can learn from them:

Let's start with a good overview of a working trebuchet from DGseward.


The raised weight represents the torque applied shoulders, the long wooden beam represents the golfers arm,

through

the

and the rope sling represents the shaft of the golf club.

Think about it; if the club had a perfectly flexible shaft -- like the rope sling of a trebuchet -- then there is no way to apply wrist torque to get any action from the clubhead. Yet the trebuchet was a very effective siege weapon. It was the best, most powerful catapult in warfare for centuries until it was replaced by gunpowder-powered cannons, demonstrating that "wrist torque" isn't all that important. Let's turn DGseward's video upside-down to show the similarity to the golf swing. It should look more like a real golf swing this way. The arms accelerate from the turning "shoulders" (the counterweight), and the club (the rope sling) lags until it is whipped outward by the tension in the rope. Trebuchet competitions remain something of a sport today. Games include distance and target competitions, with "punkin chunkin" being an American championship throwing large pumpkins for maximum distance. Here's a pretty basic "punkin chunkin" trebuchet. Finally, if you want to experiment with a trebuchet yourself, here's a science kit you could build. Golf Swing Physics 4. Technology Guest article by Rod White -- December 2008

Golf has always been a game of technology, with a long history of improvements to golf balls and clubs. However it is easy to be sceptical of the impact technology has had surely coaching and swings have improved too. On this page we look at some of the effect of some of the recent technological changes on golf driving distances.

The pink curve on the graph shows the mean driving distance for PGA tour professionals for every year between 1980 and 2008. The two curves either side show the average plus and minus 1 standard deviation (a measure of the variation amongst the golfers). The dots show the average driving distance for the longest driver of the year. The graphs show a consistent increase of about 40m with much of that occurring between 1995 and 2005. How much of that gain is due to technology? Clubhead Mass

As we should expect from the simple model we developed earlier, the driving distances are more or less independent of club head mass. As the club-head mass increases, the swing efficiency increases (more kinetic energy transferred the clubhead),

and

the

collision

efficiency

falls.

The chart shows that there is a maximum between 170 and 180 g, but it is a very broad peak. That is, you have to get a long way from the peak before it makes a lot of difference. Any club head mass between 130 g and 240 g yields the same distance within about 4 m. The optimal mass not only gives the greatest distance, it also means that the same club will do for almost all golfers (which is good news for the club manufacturers). The optimum mass indicated in the figure is still a little short of the 195 g to 205 g typical of modern drivers. There are a couple of reasons for this. Firstly, muscles work better at slower speeds, favouring a slightly heavier club and slower swing. Secondly, the driven pendulum model used to produce the graph assumes a constant shoulder torque. In reality it will take a short while for the shoulder torque to build so that more of the work is done at the end of the swing where it is not transferred efficiently, and the lower efficiency results in a heavier optimum mass. Finally, the value I chose for the effective mass of the golfers arms is a nominal value that had been used previously by other researchers, a 20% increase in arm mass is sufficient to account for the difference. Shaft Length

The effective shaft length used in the graph is defined by the point where the arms and the club hinge, and can be a couple of inches less than the full length of the club. We have already mentioned the increase in efficiency that accompanies increased shaft length. In fact, increasing the shaft length increases driving

distances through two effects: 1. An increased downswing angle which means that more work is done by the golfer, and 2. The improved efficiency of the swing a greater fraction of the work done by the golfer is transferred to the club head. The total work done by the golfer is downswing angle x shoulder torque, and clubs with a longer shaft take a longer time to swing out, and results in a greater down-swing angle, and hence greater work done by the golfer. However, a natural release swing (with no wrist torque) becomes anatomically impossible with long shafts. For the average male golfer, most of the modern standard length drivers (45-46) are a little long, although this depends a lot on the golfer's grip and the location of the centre of the hinge at the wrists. A number of professional golfers have observed that their driving accuracy improves with shorter shafts. This is probably due to a reduction in the use of the hands. Coefficient of Restitution

By far the largest contribution to increased driving distances has come about from the improved Coefficient of Restitution (CoR) of the club-ball collision. At the time of the introduction of the solid core ball, the CoR of most balls was in the low 0.7 range. As rubbers improved, the CoR improved markedly. During this time, manufacturers had to compromise between a soft skin (better spin control) and distance (with a harder, longer wearing skin). Nowadays it is possible to achieve the maximum CoR permitted under the rules while still having a soft

skin. The rules have two tests on balls, one based on ball velocity after a well defined impact with a solid steel object moving at a prescribed speed. This limits the CoR of the ball (in collision with solid steel) to about 0.79. The second limits the driving distance with a specified driving machine. With the development of metal drivers, it became possible to build large thin club faces with a resonant frequency low enough that they would flex and store energy during impact. This "trampoline effect" reduces the compressive forces on the ball, and therefore reduces the energy dissipated in the collision. As a consequence of the trampoline effect, the CoR for the combined club-ball collision has risen to about 0.83 -- where it is now limited by rule. This means, since the 1970s the improved CoR has resulted in a 20m increase in driving distances for the average golfer. Shaft Material

Another of the major technological advances has been the change from stainless steel shafts to graphite shafts.. The effect is due to an improvement in the swing efficiency. The club-ball collision time for a typical drive is below 0.0005 seconds. During this time, the kinetic energy must be carried from all parts of the clubhead to the ball. You can envisage this happening as a compression wave in the metal clubhead carrying the energy forward. The same is true for the first few inches of the shaft. However, most of the kinetic energy in the shaft cannot be carried into the head and to the ball because the speed of the wave, carrying the energy from the shaft, is too slow. So the great majority of the shaft's mass is not involved in momentum transfer to the

ball. This means that all of the work done by the golfer to put kinetic energy into the shaft is wasted -- it does not participate in energy transfer to the ball -- and reducing the mass of the shaft will make the stroke more efficient. Reducing the shaft mass from about 130 g (steel shafts) to about 60 g in modern lightweight shafts has resulted in an increased driving distance of about 10m. Note: contrary to common claims, the flex of a shaft has no effect on the club head speed, and there is little energy stored in the flex of the shaft that can be recovered at impact this is all advertising BS. The shaft flex does however affect the loft of the club at impact. The huge (50 kg for a professional) tension in the shaft pulls at the heel of the clubhead. However, the centre of mass of the club head is a few inches away from the heel, near the centre of the head, and with a high swing speed the centre of mass tends to line up with the shaft. The shaft therefore bends towards the target and toe down a few centimeters. This effect provides club manufacturers with a way of expanding the range of lofts on their clubs. A flexible shaft will enhance loft by as much as 5 degrees, while a stiff shaft only about 2 degrees. Tip-stiff shafts also reduce this effect, and tip-flexible shafts emphasize it. Golfers with a slow swing speed require drivers with a high loft (14-15 degrees say) to achieve the greatest distance, whereas golfers with a high swing speed achieve optimum distances with a lower loft (as low as 9 degrees).

Summary So there you have it! The great distances produced by good golfers using modern equipment are due to both technique and technology. The important factors we have discussed are: Technique: Use the transfer of momentum and rotational energy from your turning body and arms to the clubhead and then to the ball. Don't try to impose your will on the club, but rather allow the energy to flow to it. (This is not some newage proverb. As we saw, it is classical physics.) Particular techniques to use include:

Start the downswing with the club (and even the arms) "folded" close to the body.

Keep them folded there as late as you can into the downswing. That means the first part of the downswing is just rotation, with nothing from the hips up moving relative to any other part. Do not apply wrist torque to swing the club out or forward into the ball. You may think you are increasing the clubhead speed with "hand action", but the physics shows that it does exactly the opposite.

Focusing on the mechanical aspects of technique, as we describe it above, is often unhelpful when you are playing. Our brains work better if we can establish an emotional connection to the correct swing. When youre practicing at the range, practice by first swinging the club about your shoulders as though you are going to throw the club dont use your hands: remember how it feels, both the rhythm and the feel in your muscles. Now when you stand over the practice ball, remember the rhythm and feelings as you throw the club through the impact zone. When you hit a good shot, pause and enjoy it, and remember the feeling. Learn that feeling so you can recall it during your pre-shot routine on the course. Technology:

Clubhead speed is maximized by using the lightest, longest shaft you can control. But keep in mind the modern driver shafts are too long for the average golfer to swing without use of the hands, which means a tendency to lose control under mental pressure. Like some pros, you may find that backing off on the length and making sure of a consistent center-face impact is a better strategy for both average distance and control. Modern drivers with the larger heads and flexible face have a higher Coefficient of Restitution than older drivers, and are a little more tolerant of off center hits. Update your driver if you have not already done so. If you already have a 460 cc driver, there is probably little to be gained. Dont be tempted by drivers with a low loft. The optimum loft for modern drivers with a low centre of mass and low spin rate, is a higher than for the older drivers. Few golfers need a loft below 10.5 degrees these days. Also, as your swing improves with less use of the hands, the less the club will be ahead of the hands at impact, and the lower the trajectory. Don't worry excessively about clubhead weight. Choose it so you swing the club comfortably and consistently. Any commercially

Applying Physics to Golf

Now let's start to apply some of the physics we just learned. We'll look at what happens during the golf swing, during impact, and during the ball's flight. Nowhere near everything, of course. We will focus on some issues which are important for clubfitting, especially:

Where does the power for the swing come from? The answer will probably surprise you. What happens during impact, and how does it turn clubhead design and path into launch conditions? This is the bottom line, turning a swing into a golf shot. What determines the ball's flight? How do the launch conditions (ball speed, launch angle, and spin) translate into distance and direction. Once we have these fundamentals under our belts, let's look critically at a few myths that have perpetuated themselves in golf lore.

What powers the swing? Many golfers, especially in the United States, grew up with sports in which a ball is thrown or hit. Baseball, American football, basketball, or hockey are usually mastered -- well, at least played sort of competently -- before golf is attempted. And those sports generally involve powering the ball with the hands and arms. Yes, the body can play a part in adding to the power. But that part is generally getting the whole mass of the body moving in the direction you want the ball to travel. The golf swing is different, not just in degree but in principle. Golfers who grew up hitting things with a bat or a hockey stick have developed swing habits that are counterproductive in golf. (Well, the very best at those sports may have inorporated the important elements of a golf swing. But the average Sunday athlete has not.) Most of the power in a golf swing comes from centrifugal force, generated by the muscles that rotate the body through the swing. Before explaining it further, let's look at the physics of the golf swing. After we see the forces at work in a good golf swing, we'll go back and see what sort of bad habits most golfers carry over from their other sports. Finally, since you're probably going to be skeptical about this -- it really is counterintuitive, and you should be skeptical -- we'll review how we know this to be true. The double pendulum When an engineer sets about analyzing a real-world system -- like a golf swing -- he creates a physical "model" of the system. This is a set of elements that are simple enough to yield to calculations, yet complex

enough to represent what is actually going on. Finding the right model -- the right balance between simplicity and complexity -- is the first and often the hardest step in engineering analysis. The simplest model that makes any sense at all for the golf swing is a double pendulum. The two members of the pendulum are:

The golfer's shoulders and arms, taken as a single rigid unit. That's the green triangle in the diagram. We'll call that "the triangle" in the discussion that follows. The golf club, also taken as a single rigid unit.

The triangle is hinged to the golfer's body (the tan elements in the diagram) so it can turn. Similarly, the golf club is hinged to the other end of the triangle.

This is a very simple model, having only two moving elements hinged together. To see just how simple, let's re-draw it the way an engineer would: as a collection of free, hinged bodies. Now we can see why the model is a double pendulum; it is a black pendulum (the club) hanging from the end of a green pendulum (representing the triangle). While the diagram looks different from the golfer above, it works exactly the same when it comes to physics. Given the simplicity of the model, it's pretty amazing how close it can get to the actual measured performance of a golfer's swing. True, there are a lot of nuances of the swing that it doesn't capture. But experience has shown it is rich enough to explain where the clubhead speed comes from in a good swing.

Let's look at the next question about using the model. We have two hinges, and we can apply a torque at each of those hinges. Those two torques -- plus gravity -- are the only forces in this model that will cause the golfer to swing the club. So the engineering model has to say what kind of torque:

The body applies to the shoulders to turn the triangle. The hands and wrists apply to the club to uncock it and bring it to impact.

It turns out that the torque the body applies to the triangle is considerable, but a good swing applies almost no torque to the grip of the clubs by the hands. Yet more than half of the clubhead speed comes from the club turning about the hands at the bottom of the swing -- much more than could be explained simply by shoulder turn. What is creating that very strong rotation of the club about the hands, if the hands are not being used to supply a "hit" force?

The answer is centrifugal force. Remember that a body in motion wants to keep moving in a straight line. But the golfer is pulling the club around in a circle. According to Newton, the club wants to fly outward from the circle; the force that is trying to pull it out straight with the arms is centrifugal force. That centrifugal force is generated by pulling the club in a circle around the shoulder hinge, and the force wants to pull the club straight out along a radius from that hinge. How big is that centrifugal force? Let's look again at the formula:

= r The mass m is a property of the golf club, and the radius r is a combination of the extended arms and the wrist-cock angle. The more acute the wristcock angle, the closer the club is to the shoulder hinge -- and thus the smaller the radius. As for velocity v, it increases as the body torque accelerates the triangle. So, what does the golfer have to do to get maximum centrifugal force in order to get maximum clubhead speed? His job is to "hold the lag" -- keep the club cocked at a right angle to the arms -- until fairly late in the downswing. This keeps from releasing the club until v is nearly as large as it's going to get, which allows a large F to accelerate the club outward and downward just before impact. This, and not torque applied by the hands, is the way to reliable high clubhead speed. (It is worth noting there is criticism of centrifugal force as the mechanism of the golf swing. More on that here.) If you spent years swinging a baseball bat before you started golf, this probably defies your understanding of power. A baseball slugger has strong forearms and wrists, the better to whip the bat through the ball. Yes, body rotation is also important, but the hands are active, while for this model of the golf swing the hands are passive. On top of that, you never hear this from the pros. Instead, you hear stuff about clearing the hips, or keeping your right elbow "tucked" instead of "flying". If this is true, why don't the pros teach it? So you're probably skeptical that this is how the golf swing works. And you should be! But it really does work this way. First, let me address why the pros don't teach it; then I'll spend some time on how we know it's true. OK, why don't the pros teach it? 1. They don't know it! Very few teaching pros or TV commentators have a clue about physics. Terms like "centrifugal force" and "moment of inertia" are buzzwords they throw around without a clue what they really mean. Please don't get me started on this; it's a pet peeve and I could go on all day. 2. Actually, there are a few pros who do understand, but they are indeed few and far between -- and often skeptical. Jack Nicklaus writes about

v2 ------

an incident in 1972, when an expert in golf physics told him about this. He wasn't at all sure he believed it... "But his theory seems to explain a shot I hit at the par-3 fifteenth in the second round at Firestone. The choice of club lay between a twoiron and a three-iron, and I decided to go with an easy two-iron. Coming into the ball I was deliberately 'soft' with my hands. I've never hit a better two-iron in my life! The ball finished over the green. "Maybe this explains what happens on those good drives where I have a 'soft' feeling in my hands through the ball... My hands merely went along for the ride." 3. In a sense, they do teach it -- without knowing or believing it. The two examples above are not contradictory to the way the swing works: o "Clear the hips" creates body rotation with the large muscles, which causes rotation of "the triangle". Anything that increases torque on the triangle will contribute to power. o Tucking the right elbow has little to do with power. That move's purpose is to control the swing plane. When done right, it assures that the ball will go in the direction of the target -- and none of the analysis above deals with accuracy at all, just power. Now let's look at the reasoning why we should believe it works this way. Jorgensen's study Theodore Jorgensen set about to find a physical model that would match the behavior of a golfer with a good, "classic" swing. Here is how he went about it: He found a golfer with the sort of swing he needed. He outfitted the golfer with reflective dots on his joints, as well as on his golf club's grip, shaft, and head. He took a sequence of strobe pictures of the swing, with the reflective dots indicating exactly where all the important parts of body and club were at every moment. At this point, he had a completely instrumented swing, and could compute velocities and accelerations of club parts, body parts, wrist cock angles, etc. He started his mathematical modeling with a simple double pendulum, and fiddled with the torques until the model gave the same swing as the golfer did.

He couldn't do it with the simple double pendulum, so he added complexity a little at a time until he had an exact match between the mathematical model and the golfer's swing.

So Jorgensen's model isn't quite a simple double pendulum. The figure shows the changes he had to make to the simple model in order to get it to behave exactly like the real golfer. He has to insert a right-angle "stop" so that the wrist-cock never exceeds 90. And he had to put a little "sway" into the golfer -- a small forward motion of the shoulder hinge during the downswing. But the important change he did not have to make was to add any wrist torque to release the club at the bottom of the swing. That is accomplished completely by centrifugal force. In fact, once he had a mathematical model that behaved like the golf swing, he ran some "what if" analyses to see whether application of wrist torque could add to power. He found that there is a critical time about 70-100 milliseconds before impact (where the arms are 60 back from vertical) where torque changes from hurting clubhead speed to helping it. That is, any uncocking wrist torque before the critical time will reduce clubhead speed at impact. You can indeed increase clubhead speed a bit by applying wrist torque, but only if you can do it for just the last 70 milliseconds before impact, and not before. It takes a very well-coordinated athlete to get away with this. Interestingly, Jorgensen found that that the same critical time works the other way as well. If you use negative torque (that is, use strength in the wrist to prevent uncocking) early in the swing and then release it 100 milliseconds before impact, you will increase the clubhead speed. In fact, you'll get as much increase in clubhead speed as that well-coordinated athlete would have gotten by a late application of positive torque. And it's much easier to hold off release than to apply a release-aiding torque at exactly the right time. So Jorgensen's study confirms the notion that power in a golf swing -clubhead speed -- is a product of centrifugal force and not wrist torque. He adds a lot of detail, but nothing that denies that basic truth.

Muscle energy If classical scientific-method physics (Jorgensen's approach) doesn't satisfy you, how about biology and physiology? Coming into impact, a golf club's kinetic energy is based on its mass and speed. It gets there from zero kinetic energy during the time of the downswing, less than half a second. This implies that the muscles have to put out a certain amount of power for half a second. Physiologists know how much power a muscle can provide for a short burst (say, half a second). When this fairly simple calculation is cranked through, the answer is over 30 pounds of muscle mass needed to impart that energy to the golf club. This is muscle that is engaged in generating motion, and does not include muscle used to stabilize the body in the golf swing posture. The 30-pound number has come up consistently in quite a few separate studies aimed at this question. There isn't anywhere near that much muscle in the forearms, hands, and wrists, so they can't be the major driving force of the swing. You need the big muscles -- the legs, thighs, torso, and shoulders -- to create that much power. That verifies that the clubhead's energy comes from body rotation, not hand torque. But it doesn't unambiguously point to centrifugal force as the enabler. But we should be able to compute the clubhead speed that would result if we only used body rotation and not centrifugal force. Without any velocity at impact from uncocking the wrist, just from body rotation, we get only about two thirds the clubhead speed that a good swing actually accomplishes. So we need centrifugal force because:

We know the bulk of the power comes from body rotation. We know that body rotation without wrist-uncock velocity gives a third less clubhead speed. In order for body rotation to generate wrist-uncock velocity, we need centrifugal force -- because the small muscles in the hands and forearms can't generate that much power.

Trebuchet Still skeptical? Don't believe physics or biology? How about history... A few years ago (probably 2004), I was watching a show about Siege

Engines on The History Channel, and had a "Eureka" moment. They were talking about the Trebuchet, a rock-hurling device that was invented about 1200AD. It replaced the catapult over only a few decades, because it had more range for a smaller and lighter device. (Still big and heavy to be sure, but definitely more efficient than what came before it.) My Eureka was because, watching it, I saw an upside-down golf swing. The principle of a double pendulum driven by centrifugal force was right there, and history has proven it very effective. For a description of how a trebuchet works, see the page and the animation I clipped from The Trebuchet Store. (They sell trebuchet kits and the like, in case you find this stuff interesting for its own sake, not just what it teaches about the golf swing.) In short, the inner arm of the pendulum (corresponding to the triangle) is a rigid, pivoting structure, but the outer arm (corresponding to the golf club) is literally a string. You couldn't apply "wrist torque" to it if you wanted to -- it must operate by centrifugal force. Now one of the most interesting thing about this design is that it has never been significantly improved upon! It has been around for eight hundred years, and it is still the most efficient catapult known. Of course, catapults are no longer used for sieges; cannons and gunpowder took over a few hundred years after the trebuchet's introduction. But:

For those few hundred years, the trebuchet remained king of the siege engines. Even today, there are catapult-engineering contests; they call them "punkin chunkin'" and obviously they hurl pumpkins -- and bowling balls and other large objects up to and including major kitchen appliances. But the design that still dominates this "sport" -- even in our engineering-knowledgeable age -- is the trebuchet.

Since the equations of motion for the trebuchet are basically the same as the zero-wrist-torque golf swing, we can rest assured that the centrifugallydriven golf swing is very effective indeed. Hitters and swingers

I'd like to take this opportunity to state very specifically what I mean later in these notes by the terms hitter and swinger. Most clubfitters and many instructors make this distinction, but it tends to be intuitive and imprecise. I believe that:

A swinger is a golfer who depends exclusively on centrifugal force for clubhead speed, and adds no wrist torque during the downswing except that needed to hold a 90 wrist cock. A hitter is a golfer who depends to some extent on torque applied to the club's grip via the hands and the wrists.

Of course, there are few pure swingers and no pure hitters. But, comparing two golfers, we now have a way to say which one is more of a hitter and which more of a swinger. And, in fact, we can tell from this whether a golfer is primarily a hitter or a swinger.

Estimating Slice and Hook Dave Tutelman - November 5, 2008 This work was instigated by an email conversation with James Smith. James is a golf coach and clubfitter in Texas (Craftsmith Enterprises), and wanted to know if there was a rule of thumb to estimate the amount of curve on a deliberate fade or draw. The rule in this article may or may not be useful for teaching or course management -- or it may not. But it is physically correct and may be useful to estimate errors due to misfit clubs. We know a few things about what happens when a ball is struck with the clubface not square to the movement path of the clubhead (see the figure). 1. The ball takes off in a direction between the clubface direction and the clubhead path. 2. The ball's direction is closer to the clubface direction than the clubhead

path, in a proportion usually between 80:20 and 90:10. 3. The ball has a sidespin that increases proportionally to the ball clubhead speed and roughly proportionally to the difference angle. 4. This spin produces a draw or a fade -- or, if the spin is high enough, a hook or a slice.

Relating the difference angle to the amount of curvature (slice or hook) is a complex aerodynamics problem. Frank Schmidberger and I have developed a program (TrajectoWare Drive, which is available for download at no charge), that traces the ball's entire trajectory, including the amount of hook and slice. But you are not going to have a computer with you on the course, and that approach would not be legal anyway. So is there some rule of thumb that we can use? This was James' challenge to me. After we went back and forth a bit to better define the problem, I was able to come up with a pretty good approximation. The way I did it was to do a large number of runs on TrajectoWare Drive, then look for a simple equation that fits the data. The equation was not based on any knowledge of physics -- just any simple equation that was close enough to the data to give useful results. In other words, if I could fit a simple linear equation to within 5% of the actual data over all the runs I did, then I would not worry what Newton would say about the equation. It is usable because the difference between a 20-yard fade and a 21-yard fade can be ignored (that's a 5% difference), given the precision with which a golfer can actually produce a 20-yard fade.

Summary of Results Here is the rule of thumb. The model, shown in the figure, deals with a right-handed slice or left-handed hook. A mirror image would cover a left-handed slice or right-handed hook. So all the possibilities are covered. (Note that the figure greatly exaggerates the angles and the curvature for tutorial purposes. If such extreme curves offend your intuition, we present later a diagram that is more to scale.) In the figure, the face angle is to the right of the clubhead path, so the ball curves to the right. It is given that the golfer has a club and swing that will propel the ball some known distance. The golfer is going to hit the ball with the clubhead path X yards to the left of where the clubface is pointing, at the distance the ball will fly. The problem is to estimate the amount Y that the ball will land to the right of where the clubface is pointing. My curve-fitting study (based on drivers only) shows that the ratio of Y to X depends almost entirely on the distance that the club will hit the ball. Here are some typical values for that ratio. Distance Y/X Ratio 200 yd 225 yd 1.6 2.0

250 yd 275 yd 300 yd

2.3 2.7 3.0

So the difference between the clubface and the final angle due to slicing is even bigger than the difference between the clubface and the clubhead path. That is more than intuition tells us. I have seen lots of charts showing how much directional error is produced by a lie angle error. In every case, that error is based on the straight-line angular error. We all know that the charts underestimate the error by not adding in the hook or slice, but now we know that the actual error is more than double what the charts say. That is because there is an additional slice error that is even bigger than the straight-line error the chart tabulates. In fact, the actual error is often more than triple what the charts say. That is the basic result reported here. The rest of this article consists of:

Can it be used on the golf course? How we came up with the ratio. Is it applicable to more than just square hits with a driver?

Can it be used on the golf course? Remember that the original motivation for this study was James Smith's desire for a rule of thumb to use on the golf course when shaping shots. Here is how it might work in practice:

You want to hit a 200-yard shot with a 15-yard fade. How to you want the clubface and the swing path to be aligned to create this shot? The picture shows how the earlier diagram would look on the golf course. The ball takes off close to the face angle, so let's consider them about the same. Doing this will give us a ball that starts a bit left of the face and fades slightly more than 15 yards to make up for it. We get the final landing where we want it. (At least if we're skilled enough to get the angles right when we swing.) So we want Y to be 15 yards. Since the shot is 200 yards, X needs to be 15/1.6 (where 1.6 is the ratio that we got from the table above). So X is 10 yards. Well, 9.4 yards, not 10. That's two feet different from 10 yards. Do you think you can control your aim to 2 feet at 200 yards? If not, then round it to 10 yards. The fact that you can round it off means that you can probably do the arithmetic in your head, and be close enough to complement your ability to pull off the shot.

This strongly suggests that it should not be that difficult to use the rule of thumb on the golf course. Just set up with the club face in the initial

direction, Y from the target, and the shoulders (the swing path) X beyond that. Should work! But does it? James was not happy with this answer. His objection was that he has had success -- and successfully taught other good golfers -- to fade the ball with a rather different rule. Quoting him: The trick is to guesstimate the required amount of shoulder alignment adjustment to the left. I divide the fairway into thirds. My shoulder alignment splits the left 1/3 zone. I generally aim the clubface just inside the left 1/3 dividing line, favoring the center of the fairway. For a fade, I move the ball forward about one ball position (less than 2) from its normal position (off my left instep on most days). This encourages my hands to hold rather than fully release, reducing the tendency to close the clubface. I limit my power meter to about 90% for this situation to further reduce the tendency to close the clubface. This is not steering, it just isnt going after it full throttle. Again, I make no grip adjustment unless I really want the ball to bend around a dogleg. This technique works, but it can probably be refined to allow for a more aggressive swing. If a reasonably accurate fudge factor can be developed, then shoulder and clubface alignment can be more closely refined and taught. One of Rotellas principles is to aim at a small target. This would fit into that plan. This approach can be summarized, "Aim the shoulders where you want the ball to start out, and the clubface about 1/3 of the way from the starting direction to the desired finish direction." Assuming James' address position is duplicated at impact (that is, clubhead path where the shoulders were, and the clubface returning to where it was at address), the ball would start just inside the left 1/3 zone and finish well toward the right of the fairway, not the middle. That's not what we want. It is also not what James typically gets.

Why does James' experience differ from what TrajectoWare Drive says? After several days of emailing back and forth, James sent me this drawing explaining what the issue probably is. It goes back to, "For a fade, I move the ball forward about one ball position (less than 2) from its normal position..." This explains the difference. With the ball forward, the clubhead path will be left of the address-time shoulder line, and the clubface will have a little more time to close. So everything is a little more to the left at impact than it was at address. That explains the difference. But it is not encouraging for any rule of thumb. The aiming is very ad-hoc, depending on the golfer's making a change in ball position (move it forward) and swing (hold off closing the face). In other words, you can factor in a healthy dose of feel and minor ball position and swing changes, and make the ball follow some other rule. Nobody should find that fact surprising. But presenting the problem that way, there is no physics to any rule of thumb -- just talent, feel, and experience.. If you want to use the rule of thumb implied by the Y/X ratio, you will have to set up with the ball in the normal position and swing the way you normally do. The only difference from your normal swing is that the clubface will not be aimed in the same direction as your shoulders. Specifically, your aim keys are:

The distance between the target and the face angle is Y. The distance between the face angle and the shoulders is X. ...Where Y/X is the ratio from the table above.

Then trust your ability to make your normal "straight ball" swing.

How we came up with the ratio Taking data using a trajectory computation The results are based on output from the program TrajectoWare Drive. It is freely downloadable, so you can obtain your own copy and draw your own conclusions. A set of representative distances and difference angles were run through the program, and the results recorded. Here is a typical output from the program's "top view" window, showing the side-to-side movement of the

ball in the air. The graph axes and the "ball flight" curve are snapshots of the computer output; the rest is notation I added to link the output to the problem we are addressing.

The problem is expressed in terms of impact conditions (clubhead path and face angle), so we had to choose impact conditions for each data point. We wanted data for difference angles from 0 (straight flight) to 5 (pretty severe slice). We did this by holding the face to point at the target and varying the clubhead path in the program. Distance was a bit tricky. We wanted carry distances from 175 to 325 yards at 25-yard intervals. Distance is not a program input; it is an output based on a collection of input conditions. Here is how we chose the impact conditions to give each distance. We assumed that the ball would be hit by someone who has a well-fitted driver for their swing. The swing was chosen to have a zero angle of attack. Carry Head Dynamic Distance Speed Loft 175 200 82 90 16 14

225 250 275 300 325

97 106 114 123 133

13 11.5 10 9.5 8.5

Then we played with the clubhead speed and loft to give the desired distance. Example: We want one of our distance data points to be 250 yards. 1. Set the impact conditions in TrajectoWare Drive to 0 angle of attack, 0 face angle, and 0 clubhead path. Set typical values for the rest of the parameters (e.g.- standard ball mass, 200g clubhead mass, 0.83 COR, etc). 2. Choose a likely clubhead speed, and vary the loft (i.e.- "fit" the golfer to the proper driver) to give maximum carry. 3. Is the maximum carry more or less than 250 yards? Depending on the answer, change the clubhead speed and go back to step #2. 4. Continue iterating steps #2 and #3 until the carry distance is 250 yards. (It came out to 106mph clubhead speed and 11.5 dynamic loft.) The accompanying table shows the clubhead speed and dynamic loft for each of the distances of interest.

values

we

chose

for

Having the clubhead speed and dynamic loft for each distance, we then ran the program at each distance for a clubhead path of 0, 1, 3, and 5. The interesting output was the distance the ball was from the target line at landing -- the value "Y" in the diagram. Here is a plot of Y (the slice distance) vs the drive's carry distance, for the the difference angle.

There are several things about this graph that suggest it may be difficult to use as a rule of thumb. Most important, the three traces are curved, not straight lines. That suggests there is no simple linear proportion that would fit the data very well. The second is that the slice distance depends on both the drive distance and the difference angle.

Finding a rule of thumb in the data Is there some way to process or transform the data so it appears simpler? Yes there is. Let's remember that we set out hoping to find a ratio, Y/X, not just Y itself. If we plot Y/X for the same data, we get what looks like three nearly identical straight lines. That's much better because:

A straight line suggests a simple proportionality, which suggests a

simple rule of thumb. If all three straight lines are the same, then the difference angle is not part of the rule of thumb; it can be expressed as distance only.

It turns out that the data fits a straight line easily as well as we need it to. We can even ignore the small differences between the three lines on the graph; a single straight line will fit all three curves closely enough to be a very effective rule of thumb. The equation for the linear approximation is: Y/X = 0.014 (distance - 86yd) How well does this remarkably simple equation match the actual computer output? Here is a table so you can see for yourself. Y/X computed by TrajectoWare Drive. Linear Formula (Rule of Distance Difference angle = Thumb) 1 3 5 175 200 225 250 275 300 325 1.276 1.661 1.943 2.314 2.655 3.065 3.536 1.279 1.609 1.864 2.233 2.574 2.953 3.381 1.257 1.549 1.786 2.132 2.404 2.741 3.118 1.246 1.596 1.946 2.296 2.646 2.996 3.346

Looking at these numbers, the error is very acceptable. Remember that using the Rule on the golf course requires you to be able to divide a distance a few hundred yards away into two intervals, whose size is in the ratio of some number (say, 2.3 for 250 yards). Can you, by eye, tell the difference between dividing it 2.3 or 2.5? If not, then you'll never notice the error

shown

in

the

table.

How applicable is it? There are a couple of reasons that the rule of thumb may not be realistic except in fairly limited circumstances: 1. We have only run data using drivers matched to the golfer. What about other clubs? What about irons, which apply more spin to the ball than TrajectoWare Drive can handle? 2. In practice, a deliberately open or closed clubface is generally accompanied by a loft change. Does this affect the rule of thumb, and by how much? The short form of the answer is: The rule of thumb seems useful in both cases. Here's the reasoning that brings me to this conclusion.

Higher spin TrajectoWare Drive is currently (2008, version 1.0) limited to a spin of 4000rpm. More spin than this will give incorrect results. In fact, even the shortest hitter in the study (175yd carry, 82mph clubhead speed) will exceed 4000rpm at an 18 dynamic loft. So we are not in a position to test the rule of thumb with irons, or even lofted fairway woods. Is there some way, some reasoning, we can use to figure out whether the rule of thumb still applies? Here are a couple of approaches. (A) Try using TrajectoWare Drive a little outside its "comfort zone" The program issues a warning that it is losing accuracy when the initial ball spin exceeds 4000rpm, but it does its computation anyway. I tried it between 4000 and 5000rpm. I started with the 82mph driver who carries the ball 175 yards. Then I tried two higher-speed swings with higher-lofted clubs, adjusting the loft to still give 175 yards of carry. Here are the results.

Basic Information

Slice Information

Clubhead Dynamic Back Difference Side Y Distance Distance Speed Loft Spin Angle Spin (slice) 82 85 90 16 19 20 3600 175 4400 175 4972 175 3 3 3 661 674 708 11.6 12.3 13.0 173 173 173

What does this tell us about the Y/X ratio. If the rule of thumb applies at higher lofts, the Y column should be constant. It isn't; it shows a modest growth as the spin goes up. That modest growth is to be expected. The behavior of TrajectoWare Drive at high spin is to exaggerate the effect of spin. So we should expect more sidespin to produce more slice from the program than it would in the real world. We don't have a quantitative handle on it, but the trend is what we would expect if the rule of thumb continues to work at higher spin rates. On the other hand, the increase in Y, though modest, is at least as fast as the increase in clubhead speed. So it is possible that the rule of thumb would be better expresses in terms of clubhead speed than distance. We need more data to be able to tell. (B) Reasoning from basic principles - There are two different effects working at cross-purposes to create sideways motion: i. Higher loft needs more clubhead speed to get the same distance as lower clubhead speeds, at least at higher lofts than the driver. So we would expect more sidespin for the same difference angle, due to the higher clubhead speed. Higher loft produces less sidespin for the same difference angle.

ii.

When you look at the differences, it looks like the sidespin increase of (i) is larger than the sidepin decrease of (ii). This would lead us to expect that Y/X increases as you use more clubhead speed and more loft to get a given distance.

Bottom line - As you use more loft to get the same distance (e.g.- a tour player using a 7-iron for 175 yards of carry, compared with a senior using a driver), it is likely that Y/X increases. The interesting question is how much it increases. The increase may be small enough so that the rule of thumb still applies. It may be big enough that the Y/X ratio is more closely associated with clubhead speed than distance. We don't have enough data yet to know where the truth lies.

Collateral loft change Think about how a golfer opens or closes the clubface at address. To take your grip with the clubface in an open or closed position, you are holding the handle in a different rotational position from where it would be for a square face. If the club had a perfectly upright lie (see picture), then the rotation would just open and close the clubface. But the club has a lie angle of 5565, not a perfectly upright 90. (For what follows, 60 is close enough to any lie in the 55-65 range.) The consequence is that the loft changes as well as the clubface direction when you rotate the shaft.

If you open the clubface, you increase the loft. If you close the clubface, you decrease the loft. You get about 2 of loft change for every 3 of clubface angle change.

Does this affect Y/X? I ran several trajectories to see, using a base of 105.6mph clubhead speed, 11.4 loft, 250yd carry. Here are the results: Dynamic Carry X Y Y/X

Loft Straight shot 11.4

Distance 250 246 0 0 --

Pure hook or slice, 11.4 no loft change Slice: 13.4 3 open, 2 extra loft Hook: 3 shut, 2 less loft 9.4

12.9 28.4 2.20

241

12.6 27.7 2.20

239

12.5 27.4 2.19

Conclusion: the Y/X ratios are closer to one another than any is to the the rule of thumb (2.3). So opening or closing the face by rotating the shaft does not affect the Y/X ratio; the rule of thumb is still good.

Conclusion There is a relatively simple rule of thumb for fading or drawing a drive.

Take the distance you drive the ball, and remember a single number based on that distance, from the table. For instance, if your driver carry distance is 250 yards, the number is 2.3. Set up so that, at impact, the clubhead path, the face angle, and the target are in the relationship shown in the picture, with Y:X in the ratio of your number (e.g.- 2.3).

This simple rule may or may not be accurate enough for more lofted clubs. We do not yet have the computational tools to tell.

This is a recomputed and rewritten article. The original, from Jan 2007, was based on data from a computer program that did not give accurate results for higher clubhead speeds. Actually, the rewriting is not complete. I am gradually updating it, and things will change as I do. DaveT - Aug 5, 2009

Optimizing a Driver's Launch Parameters Dave Tutelman -- January 11, 2009

We hear a lot about how the optimum launch parameters for a driver mean you should be going for higher launch angle and lower spin. This seems to be a consequence of the sudden prevalence of launch monitors in clubfitting shops. As an engineer, I have been trying to make quantitative sense of that for some time now. I think I have finally found a way to look at it that tells me what's going on. What it tells me is that the conventional wisdom of higher launch and lower spin is an over-worrying of the situation.I'm convinced that fine-tuning the spin and launch angle is mostly a waste of time. If you get the loft right for the golfer's swing, the fine-tuning barely amounts to a hill of beans. I'll show you the numbers that lead me to this conclusion. But first let's go over what I think is the right way to fit a driver with the aid of a launch monitor. Fitting a driver Here's my technique, which has the backing of numbers that you'll see later. 1. Find the shaft and head combination that the golfer hits the best. "Hits the best" means things like: o Highest ball speed.

o o o

Best consistency of impact on the sweet spot. Consistently good swing path and clubface angle (open/closed). Good subjective feel.

Notice that I haven't said a thing yet about loft. As long as the loft is vaguely in the ball park for the golfer, you can fit the other things. They are much more important to driver success than tweaking the spin. 2. Now try out the combination with a variety of lofts. If you have interchangeable heads of the model you're going to use, that's obviously a big advantage. The goal here is: o Longest distance, as measured by the launch monitor -- or even by eye on the range, if your range is well enough marked to do it accurately. This is most likely to give the best result. o If you don't have interchangeable heads, you may have to choose a loft "by table". See how the golfer's loft compares with his/her launch angle (for this you'll have to measure the launch angle), and extrapolate to the optimum launch angle in one of the common tables. There are lots of such tables around; I just whipped up the following one, which is as good as any. Ball speed (mph) 100 115 130 145 160 175 190 Best launch angle 17 15 13 11 10 8.5 7

Be sure you do this with at least the same model head and the same model and flex of shaft. For flex, you should take measurements and not just rely on the LARSX markings. 3. Optimize launch angle and spin. On second thought, don't bother! You are not going to get more than one extra yard out of this step -two at the most -- if you did #1 and #2 right.
o

Basically, what I'm saying in #3 is: If you got the loft to give the launch angle for maximum distance for that player's ball speed, there is almost nothing left to be optimized by further tinkering with launch angle and spin. In the next section, I'll prove that. Then, in a final section, I'll look at the case for a long-drive contestant -- where every yard counts -- and we'll see what a meaningful optimization would involve. By the numbers

Note: The calculations in this study were done using the TrajectoWare Drive program, version 1.0. This program provides carry distance only, so that is what is being optimized here. As a practical matter, ground conditions vary so much that it is hard to optimize for total distance (carry + roll) with any great precision -- but we are working on a way to include it in the next version of the program. My rough and ready approach is to use about a degree less loft than the optimum, to reduce the spin and angle of descent. When I hear the words, "You should aim for higher launch angles and lower spin," I visualize the following chart:

The numbers on the chart are yardages. Moving in the direction of the green arrow -- toward higher launch angles and lower spin -- should give the fastest improvement in distance. The yellow arrows -- one with just higher launch angle and the other with just lower spin -- show more modest improvement. That is what the words imply. However, the real world is more complex. Here are the actual yardages, for a golfer with a 124mph ball speed. (That's a fairly typical male golfer with 85mph clubhead speed and consistently good impact with a "standard" driver. Note that it takes a little more clubhead speed to maintain the ball speed as the loft goes up.) Distances for spin and launch angle Spin Launch Angle (Degrees)

(rpm) 5500 5000 4500 4000 3500 3000 2500 2000 1500 1000

8 164 171 175 177 175 169 161 151 139 128

10 164 173 179 182 182 179 173 165 155 144

12 164 173 181 186 188 187 183 176 168 160

14 162 173 182 188 192 192 190 185 178 170

16 160 172 182 189 194 197 196 192 187 179

18 158 170 181 189 196 199 200 198 194 187

20 155 167 179 188 196 201 203 202 199 194

22 151 164 176 187 195 201 205 205 203 199

24 147 161 173 184 194 201 205 207 206 203

26 143 157 170 181 192 200 205 208 208 206

It's a little hard to understand the shape of this space just looking at a sea of numbers. So let's enhance it graphically. Here is a 3D picture of the yardage contour. The shape looks as if you had taken a paper rectangle and curled it along the diagonal. There is a top diagonal line going from the high-spin, lowlaunch corner to the lowspin, high-launch corner.

The two corners bent down are high launch and spin, and low launch and spin. The whole paper is tilted along the diagonal, so the highest corner is at high launch low spin. At first glance this does seem to say, "Go for high launch, go for low spin". But a closer look (below) is going to tell us to be somewhere on that high diagonal. We will see that moving off the diagonal, even if it is for higher launch or lower spin, is going to cost distance. So going for high launch and low spin helps only if both are added, and in the proper proportions to stay on the diagonal. This 3D graph shows us what is going on very intuitively. But we'll need another way of looking at the surface to get enough detail to understand in enough detail to be useful.

Here's that other way. The table below is color-coded according to the distance. The warmer the color (red is the warmest), then the longer the drive. Conversely, the purple entries are the shortest drives. This is a 2dimensional representation of the 3-dimensional surface shown above. Distances for spin and launch angle Spin (rpm) 5500 5000 4500 4000 3500 3000 Launch Angle (Degrees) 8 10 12 14 16 18 20 22 24 26

164 164 164 162 160 158 155 151 147 143 171 173 173 173 172 170 167 164 161 157 175 179 181 182 182 181 179 176 173 170 177 182 186 188 189 189 188 187 184 181 175 182 188 192 194 196 196 195 194 192 169 179 187 192 197 199 201 201 201 200

2500 2000 1500 1000

161 173 183 190 196 200 203 205 205 205 151 165 176 185 192 198 202 205 207 208 139 155 168 178 187 194 199 203 206 208 128 144 160 170 179 187 194 199 203 206

What is the "lay of the land" that we're looking at?

There is a "ridge" of maximum distance. Here's the same table with the ridge shown explicitly. It is the black dotted line, which is the maximum distance for a given launch angle. If you move along the line, the yardage changes rather slowly. If you move away from the line, you lose yardage more quickly. That's what makes it a ridge. The ridge is slanted compared to the east-west and north-south axes of the map. It goes from high-launch/low-spin to low-launch/high-spin. There is no "peak" on the chart. As you move down and to the right, the distance continues to increase -- but the increase slows to a crawl at the bottom right. (A bit more exploring with TrajectoWare Drive shows a maximum just under 210 yards, at a launch angle of 30 and a spin of 1000rpm.)

Let's re-phrase that in terms of driver performance:

As long as your launch angle and spin are "on the ridge", you are fairly close to the maximum distance you can get with your swing. If you do something to reduce (or, for that matter, increase) the spin, you must do something to also change the launch angle, or you will fall off the ridge -- that is, lose distance. Just lowering the spin by itself with not help and will probably hurt. You have change both to stay on the ridge. If you just raise the launch angle things, are a little better -- but not much. You may get a slight increase for a few degrees, before the distance goes down again. Loft Distance 8 10 12 14 15 16 18 20 149 170 183 188 189 188 181 171 Launch Spin Angle 7.5 9 10.5 12 12.8 13.5 15 16.5 1900 2400 2800 3300 3550 3800 4200 4700

So far, we're talking as if we can design a driver to place us anywhere we want in the chart. But that is not true. There is a "natural" area of the chart for drivers to live, and we have to work hard to get away from that area. Let's look at how real drivers behave. Here is the performance of a set of drivers, showing launch parameters and yardages. The table is based on a very simple set of assumptions:

86mph clubhead speed (because 85mph is not enough to keep the ball speed at 124mph, at the loft we are going to need). 200 gram head weight. Zero angle of attack; the club is at the bottom of the swing at impact.

The loft listed is the "dynamic" loft, defined at the difference between the direction the clubface is pointing and the clubhead is moving at impact. That means it includes not only the loft built into the driver but also: o Shaft bend, tilting the head up at impact. o Face roll, providing more loft as the impact point is higher on the clubface. o Cupped or bowed left wrist at impact, to the extent that this provides loft and not angle of attack. No increase nor decrease of spin due to vertical gear effect. (Here are links for an explanation of gear effect and vertical gear effect.) Very normal conditions otherwise: a ball fitting the TrajectoWare Drive model, normal temperature, altitude at sea level, etc.

The only thing we vary in this table is the loft of the driver. And we record the calculated distance, launch angle, and spin. The highlighted row is the driver with the best distance. Our 86mph golfer can get 189 yards with a 15 driver. Let's see how the collection of feasible drivers maps onto the launch space, how it fits on the ridge of maximum distance.

Since the rows of the driver table include launch angle and spin, we can plot each row as a point on the launch space table. Those points are the little golf balls. The dotted red line connecting them is the path of all drivers we could

build for our 86mph golfer by varying just the loft. Important points to note about this diagram:

Varying loft does not takes us along the ridge; it takes us almost perpendicular to it, straight across the ridge. That means that there is a fairly well-defined optimum loft. In the case of our golfer, the maximum is at 12.8 of launch angle (corresponding to 15 of loft). Not surprisingly, the maximum distance occurs where the red line meets the ridge (the black line). That's not surprising because the ridge represents the best spin for a given launch angle, or the best launch angle for a given spin. The maximum distance -- the intersection between the black and red lines -- is a long way in launch space from the maximum distance possible. It misses the peak (of 209.7 yards) by 17 worth of launch angle and 2500rpm worth of spin. If we could somehow traverse this distance of launch angle and spin, there is about 20 yards to gain -- about 10% of the distance of the drive. That is a lot, if we could somehow obtain it. (But, as we shall see, this is not easy. Moreover, it has less to do with club design and more to do with the golfer's swing.)

The lesson here is that the line of feasible drivers and the line of maximum distance are sloped in opposite directions. Where they intersect is the driver that gives the best carry. We may be able to do some fine tuning and add a few yards. But not many yards, and probably not with most golfers. Before we leave this point, here is a table giving the slopes of the two lines - the ridge and the feasible drivers -- for some representative ball speeds. Ball Speed Slope of ridge How many rpm of spin must be removed for each degree of increased launch? 140 rpm Slope of feasible drivers How many rpm of spin is added for each degree of increased loft? 187 rpm

mph

Measured at what loft? (near optimum) 18

100

mph 124 mph 150 mph 200 mph 128 rpm 232 rpm 15

125 rpm

281 rpm

12

120 rpm

370 rpm

At this point, we've worked the numbers for steps #1 and #2 of my recommended fitting procedure. That's the "lazy" use of the launch monitor. Now it is time to see just how much we can get from "fine tuning" the launch angle and spin: step #3. Fine tuning Now we know there is about 20 yards more distance, a full 10% of the drive, out there somewhere. Is it accessible, or is it the pot of gold at the end of the raibow? There is a fundamental fact of life we need to realize before we start fine tuning, if for no other reason than to set our expectations. It's really hard to raise the launch angle 17 at the same time you're lowering the spin by 2500 rpm. The loft is your strongest tool for dealing with launch angle. But loft takes spin in the wrong direction to stay on the ridge. Increase launch angle using loft, and you'll raise spin, not lower it. You have to find other methods. Such methods are few, and they generally make more modest adjustment of spin or launch angle. That said, let's look at the approaches club designers can use to increase launch while they decrease spin. Loft As we said, this story is good news, bad news:

First the good news: loft increases launch angle by almost a degree for every degree of loft. (Actually, it is not quite one-for-one, more like 80%-90%.) Now the bad news: loft also increases the spin. For our 86mph golfer with typical driver lofts, the spin increases by 270rpm for each degree of increased launch angle.

That means that we can easily increase the launch angle, but now we not only have to reduce the spin to stay on the ridge, we have to reduce hundreds or even thousands of RPM of spin just to make up for the launch angle gain. That's a high price, considering we must accomplish both higher launch and lower spin; one without the other is not much of a gain and may well be a loss. Angle of attack This is not a club design issue, but a golfer-training issue. It requires learning how to hit the ball on the upswing, which implies placing it further forward in the stance and teeing it higher. It works because you can increase the launch angle without increasing loft. Every degree of angle of attack gives a degree of launch angle with no effect on spin! How much can you gain through angle of attack? I don't have a lot of data on what golfers who try to hit up on the ball accomplish. But here are a couple of ways to estimate an answer:

Long drive data: These guys have a lot of athletic ability, plus a strong motive to learn an upwards angle of attack; they need both in order to compete. So we can look at the actual data from the high finishers in elite long drive competition and get an idea of the AoA that they manage. I plugged the launch data from two drives from the 2006 ReMax finals into TrajectoWare (they were drives Jason Zuback and Erik Lastowka), and worked backwards to the loft and angle of attack that they must have used. In both cases, the angle of attack was under 1. I was surprised it was so small. More long drive data: Tom Wishon points out that "the long drive competitors can use a head with 5-6 degs and still generate a launch angle of 11-12 degs." Even allowing for a reasonable shaft bend increasing the loft by 3, this is still 3 or 4 degrees of upwards angle of attack. Geometry: If you can somehow tee up the ball 6" farther forward in your stance, the natural arc of a 45" driver will allow hitting up on the

ball by an angle of 8. That is a lot. It will also involve teeing the ball up almost a half inch higher. This is probably an extreme upper limit to angle of attack, nowhere near available to most of us. In any event, this is not a clubfitting parameter. The job of the clubfitter is to specify the best club for the swing the golfer has. That will have some amount of angle of attack (perhaps zero or even negative), so let's move on to considerations the clubfitter can do something about. Vertical gear effect We usually think of gear effect in terms of sidespin: hooks and slices. But the same effect can work for backspin as well. Hit the ball high on the face and the clubhead rotation will reduce the nominal spin, compared to a center hit. (TrajectoWare Drive assumes a center hit when computing launch parameters from impact information.) Conversely, a low-face hit will experience increased backspin. How much can we expect from gear effect? Until a couple of years ago, most people (myself included) didn't think vertical gear effect was an issue at all. Even today, there does not seem to be much agreement on the specifics. For instance:

Tom Wishon has reported a difference of 425rpm for a 1/2" vertical difference of impact position on the clubface, with a 105mph clubhead speed. A difference of 1.5" from a very low on-face hit to a very high on-face hit allows 3/4" from a center hit to a high hit. With Wishon's numbers, that allows a little over 600rpm to be saved by vertical gear effect. Dana Upshaw has reported a difference of 3300rpm between a highface hit and a low-face hit, corresponding to a difference of 1650rpm from a center hit to a high hit. That is almost 3 times Wishon's data. Upshaw's data was at a ball speed of about 145mph, corresponding to a clubhead speed of 98mph, making the difference even more surprising.

Making use of vertical gear effect is a combination of the golfer's swing and a driver fit to that swing:

The golfer must have a repeatable impact high on the clubface. High enough to get maximum benefit from the gear effect, but still low enough not to lose COR.

The golfer can be helped a little by a low center of gravity for the clubhead. A low CG increases the range and efficacy of hitting "above the center". Shaft flex! (Finally, something we clubmakers can relate to easily.) Upshaw has reported an anecdote involving a long drive champion getting much better runout after landing after going to a flexible-tip shaft (generally considered high-launch and high-spin) and a lower loft in the clubhead. The explanation Dana gives is that the clubhead was more free to rotate and exercise vertical gear effect with the tipflexible shaft -- and I concur.

Other factors? There are other factors sometimes cited as ways to get high launch and low spin which, unfortunately, do not do what the proponents would have you believe:

Head weight distribution - A low center of gravity is reputed to give a high launch angle without adding spin. Vertical gear effect is a technical explanation for why that should be. A low CG enhances the vertical gear effect as noted above. Oh yeah. There's also the possibility of raising the launch angle by moving the center of gravity back away from the face. That definitely works to raise the launch angle. But it has the same effect as loft -increasing the spin as well -- so it's not helpful this time. Shaft flex - Stiff shafts, and especially tip-stiff shafts, have a reputation for being low spin. (Also low launch. That is not what we want right now, but remember that because it's important.) The biggest launch effect of shaft flex is on the dynamic loft of the clubface. So if you reduce spin, you also reduce the launch angle. The only valid use of shaft flex to reduce spin without lowering launch is as Dana Upshaw revealed; use a tip-flexible shaft in conjunction with a lower loft clubhead to take maximum advantage of gear effect. The shaft flex will increase the loft, the lower-loft clubhead will get the loft back down, and a high-face hit might get less backspin because the shaft will allow more clubhead rotation and the resulting gear effect. (I qualify it with "might", not "will". My analysis does not support the substantial effect that Upshaw noted. So I accept his data with considerable reservation.) Low-spin face - I don't know of a single instance of face design or face treatment that reduces spin. (At least none that is legal. Adding some sort of of slime when you play is not legal.)

Tuning up

We are not left with many tools to do fine tuning -- to try to move along the ridge and simultaneously raise the launch angle and lower the spin:

Loft moves across the ridge. Wrong direction. Angle of attack is either there in the golfer's swing or it isn't. If it is, we already accounted for it in step #2. Vertical gear effect is all that is left. Let's look closer.

First of all, the golfer must have a sufficiently repeatable swing that we can depend on high-face contact. (Maybe not all the time, but often enough to be worth fitting to.) To do such a fitting: a. Use impact tape or powder to determine ball impact. If impact is scattered, don't even to try to tune any further; this golfer needs lessons and practice, or at least attention to the gross driver fitting (step #1) that would give consistent high-face impact. Continue only if the impact is "tight". b. Find the proper tee height for high impact that still retains full ball speed. This will take a launch monitor and impact tape or powder. c. Experiment with shafts to find the one that gives the greatest spin reduction for the ideal impact. We're still monitoring impact here. Don't draw conclusions based on lower or higher impact. That will "churn" your shaft selection. If you can't get stable enough impact for this step, stop trying to fit for vertical gear effect. d. Now that the spin is lower, we have to get back to the ridge of maximum distance by adjusting the launch angle. We use loft to do this. We may need to decrease loft; that's a good thing, because it also decreases spin. Or it may be increased loft, in which case we have to spend some of our spin reduction to get back to the ridge.

Let's try an example! We'll use our golfer with the 86mph clubhead speed. We find a repeatable swing, with no discernible upwards angle of attack. The carry maxes out at 15 of loft, 189 yards of carry, and the golfer strikes the ball in the center of the clubface pretty reliably. Long drive! The place where it might pay to work your way along the ridge is long drive competition. That's a game where every yard counts. Those two yards that

were insignificant on the golf course can often make a place or two of difference in a long drive tournament. And it's OK to be inconsistent, as long as you can really crush one drive out of every 5 or 6. So let's look at the numbers there. I've plotted the spin vs launch angle space for a player with a 202mph ball speed, which corresponds to a 135mph clubhead speed. That's decent for local competition, though hardly good enough to win in Mesquite.

I haven't bothered with the contour lines, in the interest of keeping the chart simple. But I have indicated the "ridge" of the surface. This is the line we must stay on for our driver to be effective. The ridge runs from 302 yards at the low-launch end to 303 yards at the high-launch end, passing through a peak of 305 yards between 10 and 14. Also indicated is the locus of "feasible drivers", chosen with the same assumptions that we used earlier (zero angle of attack, normal conditions, etc). As we should expect, it slants across the ridge, crossing it at about 6 and 2500rpm. This is the best you can do with our simple assumptions, and gives us 303 yards. The driver that sits on the ridge has a 7 loft. This is long drive, not golf, so we're going to see if we can achieve the few extra yards available at this clubhead speed. The peak is 305 yards, and it starts at 10 and 2000rpm. (We could probably get another fraction of a yard going to 12 and 1700rpm, but we won't go for that here.) What sorts

of things can we do to increase launch angle by 4 while at the same time reducing spin by 500rpm? 1. Loft? No, that won't work. As we see from the purple line, it moves us across the ridge, not along the ridge. 2. Shaft profile? Probably won't help. Any change of launch angle due to changing the shaft is a combination of "spinless" (which works like a change in angle of attack) and "spin-inducing" (which works like a change of loft). Unfortunately, over 90% of the contribution to launch angle is spin-inducing. So it mostly works like loft, which is not helpful in moving us along the ridge. 3. Head weight distribution? I have heard that lowering center of gravity of the head will raise the launch angle. I'm prepared to believe this, and even to believe that it does it without adding spin. But I'm not certain that it works; I don't understand it in terms of the physics involved. Anyway, if it does, you can be pretty sure the manufacturers are already building most driver heads this way. So I'll assume there isn't much to be gleaned from this. Oh, yeah. There's also the possible of raising the launch angle by moving the center of gravity back away from the face. That definitely works to raise the launch angle. But it has the same effect as loft, so it's not helpful this time. 4. Low-spin face. I understand that some of the multiple-face drivers on the market (like the KZG Gemini) put a lower spin on the ball for the same launch angle. This has been demonstrated and measured. So it's a tool you can use. If you use it, you probably ought to start with it in step #1, not save it for the final tuning. 5. Higher angle of attack! This is not a club design issue, but a golfertraining issue. It requires learning how to hit the ball on the upswing, which implies placing it further forward in the stance and teeing it higher. It works because you can increase the launch angle without increasing loft. o If you have a 4 increase in angle of attack, you get a 4 increase in launch angle without adding any spin. o If you go for a 5 angle of attack and a 1.5 reduction of loft, you'll get exactly what we're looking for: a 4 increase in launch angle and a 500rpm reduction of spin. Why do I consider an angle-of-attack change feasible at all? Because longdrive competitors don't have to worry about hitting irons or wedges; they can put all their practice into the driver swing. That means they can practice hitting the ball 6" forward of the bottom of their swing, and teed 1/4" higher than normal which is needed for the 5 increase in attack angle. I would not

recommend this for a golfer, because it's too different from the swing for all the other clubs -- but it's probably OK for a long driver. Summary In closing, let me review my recommendations and the reasons for them: 1. Begin by finding the length, shaft, and clubhead that the golfer hits consistently at high ball speed. Use a loft within 2-3 of what the golfer needs, but don't worry too much about loft now -- just consistency, ball speed, and comfort. 2. Now vary the loft, using these components, to give the maximum distance for the golfer. 3. You are now within 10 yards -- 15 at the most -- of what is "on the table" for that golfer's swing. You might be able to make small changes in distance using a launch monitor and different shafts. Bigger improvement can be made if you can get the golfer to hit the ball consistently high on the clubface. Remember that errors in step #1 can cost 5-20 yards consistently, or even the ability to hit consistent drives. Don't get caught up in step #3, where fewer yards are harder won. Acknowledgements I'd like to thank Jeff Parrott of Golf Provisions (Largo, FL) for getting me serious about resolving this mystery. (Below you can see a partial transcript of the emails that got this work started.) I'd also like to thank David Bahr (from the GolfDiscussion forum) for pointing out the problems that the Wishon program had inserted into my numerical data, and Todd Kos for providing correct real-world data so I could test the various programs I use for validity. Last, but definitely not least, is Frank Schmidberger. Frank saw my study comparing existing trajectory programs, and started building a tool for his own use. Our correspondence grew into a joint venture that resulted in TrajectoWare Drive. It would have been very tedious to repeat all the calculations for this driver optimization project without our new program -so much so that this update might never have happened at all.

All about Gear Effect Dave Tutelman -- February 20, 2009

Too many questions have come up recently that require knowing about gear effect. Not just hand-waving and intuitive explanation, but the ability to estimate numbers. I can't put it off any longer; I have to do the homework. First of all, what is gear effect? When you hit a golf club anywhere but the middle of the face, the clubhead will twist. With drivers and fairway woods, the twist will impart the opposite twist to the ball. That is, if the clubhead rotates clockwise, the ball's spin will be counterclockwise -- just like two gears meshing. Sounds a bit arcane, but it is responsible for all sorts of interesting things, like adjustable weight screws that claim to cause a draw or fade, and the advice that a driver's "sweet spot" is high on the face. This article presents an analysis that allows answering questions like: 1. How much weight needs to be moved to actually induce a draw or fade (e.g.- the screw weights you see on drivers)? 2. Does shaft torque have any impact on resisting gear effect? 3. Can you really get more distance by hitting the ball high on the clubface? 4. Dana Upshaw reported that he greatly decreased a long-drive competitor's spin by going to a flexible tip and lower loft. Vertical gear effect is the only way I could imagine to get the sort of results that Dana got, but it requires that the shaft's tip stiffness be a major factor in how much gear effect there is. So how does shaft stiffness influence vertical gear effect?

5. I am updating my article on optimizing the launch parameters for real drivers. It requires knowing whether vertical gear effect provides enough spin reduction to make a significant difference in distance. All these questions require a quantitative knowledge about gear effect. Not precise knowledge perhaps, but at least good ballpark estimates. This article, after a tutorial introduction to gear effect, presents the results of analysis in a summary form. The subsequent pages present the analysis itself, and detailed answers to the interesting questions, along with more diagrams and photos. Introduction to gear effect Let's start with a brief explanation of what gear effect is. The explanation is lifted from my tutorial on golf club physics. Gear effect is sidespin which is the result of an off-center hit with a club whose center of gravity is well back from the clubface. Without both these conditions, gear effect does not happen.

Let's see what causes gear effect. In the picture at the right, we have two off-center impacts, one on an iron and the other on a driver. Both are toe impacts, which means it is to the toe side of the center of gravity of the clubhead. (The CG is denoted by the fourquadrant black-and-white circle; it's a pretty common notation for CG.) What does Newton say about such an impact? The CG wants to continue moving forward in a straight line, but there is a force on the clubhead that is off that line. That creates a torque that wants to twist the club. The result is that the CG keeps moving forward, but the club rotates around the CG in

clockwise

direction

(red

arrows).

The CG of the iron is close to the clubface. So, where the clubface and ball meet, this rotation (the red arrow) consists of the clubface "falling away" from the ball. This results in loss of distance (the momentum transfer is not as complete as it should have been), and perhaps the ball flying somwhat to the right as the face opens. But there isn't any special effect on spin. The driver is a completely different story. Its CG is well behind the clubface. When the driver head rotates around its CG, the whole face of the club moves sideways. Look at the direction of the red arrow where the clubface and ball meet; it is mostly parallel to the clubface, with only a bit of "falling away". So the club's face is moving to the right while the ball is compressed on it. The result is that the ball starts to rotate so its surface doesn't slide along the clubface; remember it's compressed so sliding is difficult. This rotation is

the blue arrow in the picture. If the clubhead is rotating clockwise (as in the picture), then the ball rotates counter-clockwise. It's as if the clubhead and ball were a pair of gears, with their teeth meshing where they meet. That's why a toe hit with a driver tends to hook. For all the same reasons, a heel hit with a driver tends to slice. You don't have this effect with an iron. Summary of results Here is an executive summary of the results. Click on the topic header to see more detail than the summary provides. Horizontal gear effect The picture shows the model we use for the analysis.

The impact is a distance x from being through the center of gravity (CG). The CG is a distance C behind the face. The ball leaves the clubface with a velocity Vb. The moment of inertia of the clubhead about its CG is Ih.

With C and x in inches, Vb in miles per hour, and Ih in gram-cm2, the sidespin due to gear effect in RPM is given by: Vb C x s = 58,830 Ih After analyzing how C and Ih vary in current driver heads, the equation can be further simplified to: s = 16.4 Vb x with pretty good accuracy for the vast majority of designs.

An interesting by-product of the analysis, worth noting and using, is that the force (in pounds) between ball and clubhead is: F = 9.24 Vb Here is a graph of the hook or slice spin due to gear effect, for values of ball speed Vb from 80mph to 200mph. (If you prefer to think in terms of clubhead speed, here is a conversion table.) If you use this spin to look at the trajectory, you must also take into account the face bulge radius. For instance, a toe hit will provide hook spin via gear effect, but bulge will cause the ball to start to the right, and to have some slice spin that will subtract from the gear effect hook spin. Vertical Gear Effect If you strike the ball above or below the center of gravity, gear effect will occur in a vertical direction. Let us start with the assumption that there is no reason to expect this to be substantially different from horizontal gear effect. So, if the vertical miss is y and the moment of inertia in the vertical plane is Iv, then the gear effect spin is given by: Vb C y s = 58,830 Iv Some approximate calculations suggest that, for most driver heads, Iv will

be some fraction of Ih, probably between .5 and .66 of Ih. Consequently, the spin due to vertical gear effect is between 1.5 and 2 times the spin due to horizontal gear effect, for the same amount of miss. (Of course, there isn't as much room on the clubface to miss vertically as horizontally.) As with horizontal gear effect, we can approximate the spin for most drivers. The vertical approximation is: s = 25 Vb y Vertical gear effect is the major reason that golfers are told that the "sweet spot" of a driver is above the center of the clubface. Better golfers (and perhaps all golfers) will get more total distance from a higher launch angle at the same time as lower backspin. Vertical gear effect can reduce backspin without reducing launch angle -- in fact, it may even be accompanied by an increase in launch angle. The reason for reduced backspin is that the gear effect from a high-face hit will produce topspin. The ball does not experience a net topspin; the backspin due to loft is much too great for this. But the topspin is subtracted from the backspin, reducing the backspin. Assuming a properly fit driver, the result is increased carry distance as well as increased roll after landing. A sample calculation shows, for a 150mph ball speed, a gain of 8 yards of carry and a reduction in angle of descent of 6 for a hit 0.6" above the center, compared with a hit in the center of the clubface. The only big surprise here is how big the spin due to vertical gear effect is. Strikes that are extremely high and low on the clubface (but still on the face) can result in 1500-3000rpm of gear effect spin -- much more than most people believe. So which is true, the model or what some people believe? The model has been validated using data from Hotstix published in Golf Magazine.

Application: Do weight screws produce draws and fades? We see drivers on the market today with weight screws that claim to control the trajectory of the drive. The TaylorMade R7 family was the original driver to offer weight screws, and is still the best known. The latest R7 is advertised to be able to make a 35-yard difference between maximum fade and maximum draw, just by placement of the weight screws. (The screws

allow moving 15 grams between the heel and the toe, the biggest weight shift of any R7 model so far.) Does this claim have any validity? For years, I have been saying it does not - that weight screws do not move the center of gravity of the clubhead enough to produce a noticeable difference in performance. I was not alone in this view; Tom Wishon has been most vocal that you need 30-40 grams of "discretionary weight" to see the difference. Now we have a mathematical model that allows us to predict the effect of moving the weight. What does it tell us? It turns out that Tom and I were wrong. There is a noticeable draw/fade effect from moving even as small a weight as 15 grams. Here are my conclusions:

The size of the draw/fade varies a lot with ball speed. Not only does the spin vary with ball speed (we know that from the equation for gear effect spin), but also the time and distance the ball is in the air letting the shot curve. In order to achieve the 35 yards that TaylorMade advertises, you need a clubhead speed about the average of the Tour players -- 115mph. What about the likely target of the advertising: the hacker who hits the ball with an 85mph clubhead speed and usually hits a high, 40yard slice. Unfortunately, this golfer will only see a 6-yard improvement to that slice, and a golfer like that is unlikely to even notice such a small improvement relative to the slice that remains.

Application: Does shaft torque limit the gear effect? It turns out to be fairly easy to show that even very torque-stiff shafts have negligible limiting influence on gear effect. The reason is that impact lasts such a short time that the clubhead only has an opportunity to rotate about 2 during that time. Even with a very "low torque" shaft with a 2 rating, that only produces one foot-pound of torque in the shaft. Meanwhile, the ball is imposing more than 80 times that torque on the clubhead. So the inertia of the clubhead is absorbing about 99% of the ball's torque, and the shaft only about 1%. Therefore the difference between that low-torque shaft and a torsionally very "loose" shaft is going to make a 1% or less difference in the gear effect spin.

Application: Does shaft tip stiffness limit vertical gear effect? This question is harder to analyze, and the answer I got is more equivocal and less satisfying. The tip stiffness, according to the analysis, may have a measurable effect on the spin -- but a far smaller effect than Dana Upshaw's anecdotes report. Dana shows a difference of thousands of rpm as tip stiffness is varied. My analysis can only account for a few hundreds of rpm. The limitation due to tip stiffness seems to max out at less than 14% of the spin. That is more significant than the 1% we got for shaft torque, but still not a really big deal.

Application: What is the role of roll?

Most drivers not only have bulge but roll as well. That's curvature of the clubface in the vertical plane. Does face roll play some important role in the flight of the golf ball? If so, it is a help or a hindrance? The graph shows that roll is a considerable help. Both drivers give the best distance when impact is about a half inch above the center of the clubface. (The reason for that is vertical gear effect.) Away from this best height, carry distance falls off. But, as the graph shows, a driver with the best possible face roll will not lose distance nearly as fast as a flat face. The blue curve has a very sharp peak, and loses distance rapidly as you move away from the "sweet spot".

According to the mathematical model, the optimum face roll for drivers is about an 8" radius -- a little flatter for slower clubhead speeds and a little rounder for big hitters. Even if the model's assumptions are off, I think it's unlikely that the optimum roll is more than 12", and probably less. So the next time one of the TV golf gurus says, "Today's drivers are designed to be hit high on the face," you can answer, "Yeah! They put loft on them." Seriously! Gear effect has been a fact of life since golf has been played. And a driver will give maximum distance when hit high on the face because of gear effect and loft. That is nothing new! What is new is that, with fitting and training being done with launch monitors, we suddenly know that we get better launch conditions from a high-face hit. And now you know why -- and those TV pundits do not. So what does this say about Graduated Roll Technology (GRT)? GRT is Tom Wishon's clubface design with greatly reduced face roll. The analytical model certainly does not seem to support the concept.

This pair of graphs shows the computational model's output when comparing the optimum roll with the roll numbers of GRT. Not only does GRT incur a carry distance penalty for low-face hits, those hits produce a considerably

higher angle of descent. Angle of descent is a strong indicator of distance after landing, so the the penalty in total distance will be greater than just the carry distance penalty. This agrees with my limited personal experience with GRT. It does not agree with what Wishon says his prototype testing showed.

Conclusion Gear effect, both horizontal and vertical, is remarkably important for club designers and custom fitters. The club or component designer needs to know about it in detail. The fitter needs to know about it in principle, in order to allow the golfer to take advantage of it -- or perhaps minimize it for the golfer who cannot use it to advantage. Acknowledgements I'd like to express special thanks to Russ Ryden, whose high-speed videography was instrumental in figuring out the effect of shaft tip stiffness on vertical rotation of the clubhead. Thanks are also due to Richard Kempton and email discussions helped focus my thinking discussion turned to face roll and GRT, I got from Ed Reeder, Malcolm Shepherd, and especially Jeff Summitt, whose about the problem. When the some very helpful suggestions Alan Brooks. Thanks, guys!

Notes: Ball speed, clubhead speed, and smash factor Ball Speed Smash Factor

Clubhead 1.48 1.40 1.25 speed (pro) (decent) (hacker) 70 80 90 100 110 120 104 118 133 148 163 178 98 112 126 140 154 168 88 100 113 125 138 150

The equations for spin are in terms of ball speed. But most people tend to think in terms of clubhead speed. Here is a conversion table based on the simple equation: BallSpeed = SmashFactor * ClubheadSpeed The "smash factor" is a number based on clubhead characteristics (like COR and mass) and on the goodness of the ball strike. A clean, on-center, square-face strike with a modern driver has a theoretical maximum smash factor of 1.5. The tour pros typically have a smash factor in the mid to high one-point-fours, say 1.48. Fairly decent golfers are in the low one-pointfours, and hackers can be 1.25 or even less. Note that the 1.50 maximum smash factor assumes a 200g clubhead (typical for a driver), a 0.83 COR (the legal maximum), and almost no loft (loft contributes to "unsquareness" of the hit). The Nine Ball Flights Dave Tutelman - June 3, 2006

On May 25, 2006, Ed Johnson posted owner/moderator of the Spinetalk forum):

on

Spinetalk

(Ed

is

the

I know there are 9 basic ball flights in golf shots. Straight with square face, open face, and closed face, Right with square face, open face, and closed face, and Left with square face, open face, and closed face. I've seen graphics depicting these ball flights in many places but now that I need one, I can't find one! Anybody have a link to a graphic described above? A link was posted to the "standard" diagram (see picture at left), and Ed was satisfied. But this standard picture does not show the nine trajectories that Ed asked for. Ed asked for:

Straight with:

square face open face closed face square face open face closed face square face open face closed face

Right with:

Left with:

Giving the graphic its due, the flight patterns are correctly named, and the influences on ball flight are correctly described. But unfortunately, the names are not related to cause and effect, and there is no correspondence to the nine trajectories that Ed asked for. In fact, Ed's question is ambiguous. In order to understand why, we need to ask, "What does 'square' mean?" There are two equally valid answers:

1. Square to the target line, and 2. Square to the clubhead path (the direction the clubhead is traveling). Unless we're talking about a straight down-the-line clubhead path, these two definitions of "square" give very different clubface positions. And the ball flights will reflect these differences. By the way, the nine trajectories shown above do not answer Ed's question for either definition of "square". So let's go back to basics, and get a set of ball flights that correspond to Ed's question. Actually, we'll need two sets of ball flights: one for each interpretation of what "square" means. The diagram at the right shows how the ball comes off the clubface if the clubhead is not moving in the same direction that it is facing. Here are the basics of what happens:

The ball will take a direction (red arrow) somewhere between the direction the clubface is pointing and the direction the clubhead is moving. The ball's path will be closer to the clubface direction than to the swing path. Most references show this as between 80:20 and 70:30. That is, the ball is 80% of the way from the swing path to the clubface direction.

The other obvious consequence of the clubface direction being different from the swing path is spin. The conditions in the diagram will result in clockwise spin on the ball, resulting in a fade or slice. How does this relate to the "usual" diagram shown above. Well, it would relate very well -- if only the direction of the ball were well aligned to the swing path. But it's not; instead, the direction of the ball is closely aligned to the clubface direction. This fact is not reflected in the "usual" diagram.

When we take this inconvenient fact into account, we get a somewhat different set of ball flights on our diagram. Below are two diagrams. Each has nine trajectories on it, corresponding to the nine ballflights that Ed asked about. In the diagrams:

Red arrows correspond to an outside-to-in swing for a right-handed golfer -- that is, a swing path to the left. Green arrows correspond to a down-the-line swingpath, straight at the target. Blue arrows correspond to an inside-to-out swing for a right-handed golfer -- that is, a swing path to the right.

This graphic shows ball flights where the clubface direction is referenced to the target line. That is, instead of using the ambiguous term "open", we say the clubface points right of the target line. The different kinds of arrows mean:

This graphic shows ball flights where the clubface direction is referenced to the swing path. That is, instead of using the ambiguous term "open", we say the clubface points right of the swing path. The different kinds of arrows mean:

Solid arrow: clubface points at the target. Dashed arrow: clubface points right of the target. Dotted arrow: clubface points left of the target.

Solid arrow: clubface points the same direction as the clubhead travels. Dashed arrow: clubface points right of the swing path. Dotted arrow: clubface points left of the swing path.

A few notes about the graphics: 1. If you click on the pictures above, you will get a larger-size standalone picture suitable for printing. 2. The angles for both the swing path and clubface directions are the same. That is, the swing path is 19 right or left, and the clubface is facing the same 19 right or left (of the target or the path). The curvature of the extreme slices and hooks are not quite to scale; there wasn't room on the page. 3. A consequence of this is that there are actually a lot more possible ball flights than can be shown here. For instance, the usual diagram's trajectory #3 is a pull-slice that starts out hard left and slices back to the middle or even the right of the fairway. This -- frankly very common -- ball flight comes from: o A clubface pointing well left of the target (causing the ball to start to the left), and o A swing path even further to the left (because it has to be even further left to put a slice spin on the ball).

A discussion on the Wishon Golf web forum criticized these findings. In particular, Bill (a professional clubfitter from Santa Barbara) argued that "ball flight rules" said that the ball started in the direction of the swing path and curved toward the direction of the clubface. Simple -- but wrong. (H.L. Mencken once said, "For every problem, there is a solution that is simple, neat, and wrong.") Here are my responses to his arguments: 1. As a rebuttal to Bill's version of the "ball flight rules", I pointed out that, whatever they are, they must work in the vertical direction as well as horizontal. If they were as Bill proposes, a wedge shot should start out horizontal and climb in trajectory only due to spin. But we have all seen personally that wedge shots take off on a rather high trajectory, closer to the loft angle than to horizontal. So it is just wrong to say the ball starts off in the direction of the clubhead path; we're only arguing about how close the ball starts to the clubface angle. Which brings us to... 2. Bill pointed out that 19 is a huge amount to be off, either clubface angle or swing path. And so it is for "misses"; it might not be for a deliberate hook or slice. But sorry, Bill; for smaller angles, the results are pretty similar -- but more so. That is, instead of the direction of the ball being 70-80% in the direction of the clubface, it will be more like 90%. 3. If you are still unconvinced, here are some slow-motion videos of impact. The clubhead path and clubface angle were deliberately chosen to show the true ball flight rules. (And the side comments about the old rules are downright sarcastic. Sorta' what I'd like to say, but I'm too polite.) Bottom line: I stand by what I say here. Application: Diagnose a Slice This is not just an academic exercise. I find it shocking how many instructors -- even the supposed swing "experts" on TV -- give lip service to the new rules but don't have a clue about their implications for debugging a swing. The most common example of this is the misinformation around about the cause of a slice. Sure, there is general agreement on the indisputable facts that: 1. A slice is caused by the clubface being pointed to the right of the clubhead path (for a right-handed golfer), creating spin to curve the ball to the right.

2. It can be due to two kinds of swing flaws: an open clubface or an outside-to-inside swing path. But, when it comes down to the nitty gritty of diagnosing a particular slice, most of the golf books and instructors ignore the ball flight rules and go by old wives tales. Here are the two most common slices:

The straight slice starts out straight at the target, then curves disappointingly to the right. The pull fade starts out at the right rough, and curves back to the target line. (Ever note how a good result gets called a "fade" and a bad result gets called a "slice"? No, I'm not going to comment on that.)

Let's see if we can diagnose the cause of these two slices. Most of the literature around gets it wrong, as do a lot of instructors. Diagnosing the straight slice Here we have a shot that starts out at the target and curves well to the right. Most teaching says, "Oh, you have an open clubface at impact. If you can square up the clubface, you'll be fine." Let's see what the ball flight laws would say.

Remember that the ball starts out very close to the direction the clubface is pointing. If the ball starts out at the target, then that is where the clubface was pointing. No, you do not have a problem squaring up the clubface. In order to generate slice spin with a square clubface, the clubhead path must be to the left, an out-to-in swing. This is going to be harder to fix than originally thought, because a square clubface is generally easier to achieve than a square swing plane. The trajectories for this section were generated using the computer program TrajectoWare Drive. The specifics for this straight slice were: Clubhead path = 3 left Clubface to target = 0 Resulting angle clubface to path = 3 right The ball starts 1yd left of the target line (essentially on target), and finishes 18yd right. Diagnosing the pull fade Most instructors look at a pull fade like this one and say, "You're coming over the top, but your clubface is pointing nicely to the target." Closer than before, but still no cigar.

If the ball starts left, then the clubface must have been pointing left. If it then curves right, the clubhead path has to be even more to the left. So you have a shut face but an even more severe out-to-in swing path. This consists of two errors that need correction, including a severely overthe-top move. And you need to fix both! If you fix either one by itself, the golf results will get much worse. (A pull hook if you get the swing plane fixed, and a really big slice if you get the clubface pointing at the target.) My recommendation (unless the golfer wants to spend a lot of time with lessons and practice) is to live with the good result and not worry how the ball got there. Bruce Lietzke made a fine career on the PGA Tour by not messing with this, his well-grooved swing. And, as they say, "There's no room on the scorecard for pictures." The computer specifics were: Clubhead path = 8 Clubface to target = 5 Resulting angle clubface to path = 3 The ball starts 19yd left of the target line, finishes 1yd right Diagnosing the push slice Let's look at one more scenario, the push slice. This one starts somewhat right and curves much righter, often eliciting an "Omigod!" from the golfer. It's so horrible you imagine it is the result of lots of very wrong things, and would be hard to fix if chronic. Nothing could be further from the truth; in fact, it's so easy to fix that it can be done during a round of golf. (There are left left right

very few things I would advise a golfer to mess with during a round without practicing it first, but this is one of them.)

Of course, before we fix it we must diagnose the cause. And this one is easy. It's a wide-open clubface, pure and simple. The ball starts right because that's the way the face points. If the swing path is down-the-line, then that adds slicing curve. And there you have it: a good swing path and an open clubface. I have often helped golfers fix this on the course. When I see them pushslicing routinely, I have them address the ball and ask them, "How many knuckles do you see on your left hand?" I have them strengthen their grip so they can see more knuckles. It's a quick fix that requires no swing change and works every time. It pays to know the ball flight laws! The computer specifics were: Clubhead path Clubface to target Resulting angle clubface to The ball starts 9yd right, finishes 30yd right = path = 3 = 3 0 right right

Note that difference examples, generated

all three examples had the same curvature to the shot; the was where the shot started out. That's because, for all three the clubface was pointed 3 to the right of the clubhead path. I the examples by picking a starting direction, aiming the clubface

there, then picking a swing path 3 to the left of the clubface direction. Works every time.

Bibliography and acknowledgements Where does my information come from? Here are a few reference materials that collectively add up to the story as I tell it above:

Theodore Jorgensen, "The Physics of Golf", American Institute of Physics Press, 1996. Jorgensen, a Professor Emeritus of Physics at the University of Nebraska, undertook several studies -- experimental and analytical -- to decipher the mysteries of golf mechanics. One chapter is devoted to the direction the ball leaves the clubface. Cochran & Stobbs, "The Search for the Perfect Swing", The Golf Society of Great Britain, 1968 (reprinted many times since; my copy is the 1994 printing). This is the best beginner's book on golf physics. Chapter 23 is entitled "The Ballistics of Golf: How Spin and Flight Begin". Tom Wishon's "Trajectory Profiler" program (version 2.0, 2005) incorporates the latest thinking on golf ball ballistics and aerodynamics. Since physics works vertically as well as horizontally, you can use loft to simulate an open or closed clubface and determine the resulting initial ball direction and spin.

I'd like to thank Ed Reeder for a few suggestions that made the graphic better.

The Double-Pendulum Model and the Right Arm Dave Tutelman -- November 30, 2010

Many instructors and some golfers criticize the double-pendulum model of the golf swing as inadequate. The most frequent complaint is that it fails to reflect the role of the right arm (in the right-handed swing). Interestingly, physicists and engineers seldom offer this criticism, because they know how to incorporate the effect of the right arm. Here's what they know that the instructors don't. This article covers what the model is and what it isn't, what it can tell you and what it can't.

The Model

The doublependulum model is the basis for every physical analysis and simulation of the golf swing that I have seen. It holds that the moving parts of the swing are an inner member (green) representing the arms and shoulders, and an outer member (black) representing the club. The green member rotates around a hinge at the top of the spine, and

carries a hinge itself (the wrists), to which is fastened the black member. Both hinges may be driven by torques; most analyses call these the shoulder torque and the wrist torque. These torques can be positive (counterclockwise, causing the swing to evolve in the usual direction), negative (clockwise, retarding the evolution of the swing), or even zero. In fact, the swing many physicists refer to as the "standard swing" uses a zero wrist torque for most of the downswing. (For a refresher course on torque, click here.) It is probably important to mention that the model -- the double pendulum - is the picture on the left. It is two rigid "sticks" connected by hinges, which may be powered at the hinges. The picture on the right is redrawn so a golfer will recognize it more easily. But don't attribute any anatomical detail to the way it is drawn. Ultimately:

The tan body (legs, hips, torso, and head) are really just a stand that can exert a torque on a hinge pin. Other than that, the entire body is rigid. The green triangle (shoulders and arms) are really just a rod hinged at each end. The body can exert torque on the rod at one end, and the rod can exert torque (or not) on the club at the other end.

So it is probably safer to think about the model than the nice drawing that reminds you what's what. You don't want to get too attached to the idea that you can see arms or legs in the drawing; they are not part of the model.

What we're torquing about Here's another critical place where you might not understand the model. Are you sure you know what we mean by "wrist torque"? We better make sure. Too many golfers, even instructors, hear the word "torque" and immediately assume we are talking about twisting the shaft around its axis. That is not necessarily what an engineer or physicist means by torque. To someone who's familiar with torque (which is covered in my physics tutorial), it is a turning force. The important points are:

The force causes the object it acts on to turn, rather than just move. The magnitude of the torque is the force times the distance between the action of the force and the axis of the turning.

The picture shows what we really mean by wrist torque, when we use the double pendulum model. It is not a force that tries to twist the shaft about its axis, but rather a force from the hands that tries to turn the club down to the ball -a force that releases the wrist cock.

There are certainly many variations using this model. For instance, the "standard swing" applies a positive wrist torque early in the downswing, for the purpose of keeping the wrist cock angle from falling into the middle of the swing. This torque is usually represented by a "stop" that holds the wrist cock angle; but it is really a torque that peaks very early and drops to zero about a tenth of a second into the downswing. By that time, centrifugal force on the club is more than enough to keep the wrist cock angle from shrinking; in fact, the club wants to fly out -- to release. So no further wrist torque is needed for the stop. This model is simple enough to set up differential equations that can be simulated pretty easily on a computer. In fact, it doesn't have to be a workstation of the class used to design clubheads using Finite Element methods. Any PC can run Max Dupilka's SwingPerfect program at visually instantaneous speed. But is the model useful? Does it model enough detail of the swing so the results of the model are not just an academic exercise? I think so. Most physicists and engineers feel the double pendulum represents reality well enough to:

Evaluate changes in club construction -- at least those coarse enough to be characterized as mass positioned on a rigid body. These include club length, the masses of the shaft, clubhead, and grip, and the shaft

balance point. Evaluate some advice on how to swing, or even to train for golf; i.e.how much clubhead speed comes from increased shoulder torque or wrist torque.

In the final analysis, we have to remember that it is a model. We don't expect it to be reality, but it has to reflect reality at some level. The level may not be useful for instruction -- often not even close. But the model is useful to evaluate some kinds of swing changes, to find their effect on clubhead speed and loft at impact. What kind of swing changes? Those that can ultimately be described in terms of shoulder torque and wrist torque. Many instructors feel the model is inadequate to analyze their favorite detail of the swing. In some cases they are correct. But the most commonly cited shortcoming is the folding and unfolding of the right arm -- and the model has a very effective way to deal with that. Let's see how we can deal with complex right arm action in this very simple model. The Right Arm

Most instructors feel that the folding/unfolding of the right arm (or pointing/unpointing, as one put it recently) is an important part of the swing that the double pendulum fails to model. Yes, it is an important part of the swing. But the double pendulum is quite capable of modeling the important thing it does -- add forces to the grip with the right hand. One instructor actually insisted that I change my analysis to the model in this picture, changing the two hinges of the double pendulum model to six powered hinges. Honestly!

Fortunately, the real effect of the right arm produces something the model was designed to handle: wrist torque. Let's look at this in more detail.

The right arm and wrist torque

The point of the analysis is to compute the progress of the swing, in terms of club position and speed. So any additional complexity due to the right arm boils down to what the right hand does to the club. In three dimensions, it can do six things to the club: 1. Force in the swing plane -- the red arrow in the picture. 2. Torque in the swing plane -- the circular blue arrow in the picture. 3. Force perpendicular to the swing plane. (If we could show it, it would be towards or away from the viewer of the picture.) 4. Torque perpendicular to the swing plane. 5. Axial force -- force along the axis of the shaft. 6. Axial torque -- rotation of the club around the axis of the shaft. All of these forces can theoretically be exerted by the right hand on the grip of the club.

But let's remember what a double pendulum model is used for. It computes the progress of the swing, in the swing plane. It is a two-dimensional analysis.

Motion of the club (or golfer) perpendicular to the swing plane is not interesting to the analysis, nor is axial rotation of the club. Both are very interesting to the direction the ball will travel, but you'll never see a double-pendulum analysis that deals with clubface direction, except for wrist cupping or bowing that adds or subtracts loft at impact. Most issues of clubface direction and clubhead path require a totally different approach to analysis. As for axial force, it turns out to be the way shoulder torque is conveyed to the club. It has already been accounted for in the shoulder torque.

So, from the entire list above, only #1 and #2 (in-plane force and torque) are relevant to a double pendulum model's job. It is easy to see how the inplane torque is accounted for; it is simply part of the wrist torque. But we still don't see where a double pendulum is able to account for the force. This picture shows that the model can also handle the in-plane force. Any

force exerted by the right arm in-plane is a push or a pull on the grip. (The picture shows a push.) That force works against the pull or push of the left arm. In other words, the left hand acts as a fulcrum or pivot, and the right hand's force tries to turn the club around this pivot. But what is such a turning force? It is a torque. And how big is it? It is the size of the force, times the distance between the force and the fulcrum. So any action of the right arm can be factored into a double pendulum model as wrist torque, because either it is a torque (the blue arrow above) or it can be computed as one (the red arrow above). Here's a bonus observation:

If your favorite swing uses wrist torque for either control or power, then a ten-finger grip (or baseball grip) is appropriate. That's because you increase the distance between force and fulcrum, thus increasing the wrist torque. If, on the other hand, you believe the "standard swing" model is right for you (with no wrist torque), then an overlap grip is preferable. The less distance between the hands, the less opportunity for unwanted hand action to produce wrist torque. In fact, some good players (Jim Furyk comes to mind) use a double-overlap grip, which further reduces the torque-multiplying separation between force and fulcrum.

The right arm and shoulder torque

For a long time, I didn't think the right arm could affect shoulder torque. But some recent analysis suggests that it can. No, it doesn't actually affect the shoulder torque itself. But you can model certain right arm actions as a change in shoulder torque. And, since the double pendulum is not reality but rather a model useful for analysis, that is as relevant as anything could possibly be. The diagram at the right is a re-drawing of the double-pendulum model. The important thing here is that the length of the inner member of the pendulum (representing the arms) defines a circular arc of radius R, and the hands move along that arc as if they were on a track. They are driven around that track by the shoulder torque. Expressed more accurately...

They are driven around the track by a force due to the shoulder torque, shown in the diagram to the left. Since the path is a perfect circle, the magnitude of the force is Force = ShoulderTorque / R and the force acts exactly tangent to the arc.

Anybody familiar with physical modeling can see immediately that both pictures of the model -- the double pendulum and a circular track for the hands -- gives an identical result for any analysis. So, even though it looks different and the equations you get might look different, the answers will come out the same. Anyplace you could use a double pendulum analysis, you could use a curved-track analysis. Why is this even interesting? Because one of the criticism's leveled against the double pendulum model that the path of the hands is not circular. Due to the folding of the right arm, the radius may be shorter at the top of the backswing. (May be. But not "must be". It depends upon which muscles transfer the shoulder torque to the hands, which is an issue of swing keys and technique.) Does this invalidate the model? Probably not. For years, it has given realistic, accurate results when modeling the swings of real golfers. Cochran

& Stobbs noticed this as early as 1968. Jorgensen quantified it in the mid1990s. Other researchers have been similarly successful using it to model the real golf swing. But sometimes it is necessary to take into account the folding of the right arm. When that happens, the circular track becomes a useful modeling tool, more useful than the original pendulum representation. If the path does not vary too radically from a circular curve, you can use the diagram below to represent the hands as a carriage moving along a track.

In the diagram:

The dotted line shows the direction of the center of rotation (the shoulder pivot). The center is a radius r from the hands. (The radius is lower-case this time, to indicate that it varies over the course of the swing.) The black dashed line is the actual path of the hands. The important thing to notice here is that it is not a circle around the shoulder pivot; if it were, then it would be perpendicular to the dotted-line radius. So there is some angle between the actual path and an ideal circular path. Shoulder torque produces a force along the ideal circular path; that is what torque does. The force is shown in blue, and is of a magnitude equal to the torque divided by r. The force can be resolved into components, shown in aqua. One component (the bold one) is parallel to the path of the hands, and so accelerates the hands along their path. The other (the very pale one) does nothing to aid or hinder the progress of the hands, and is rather

small as well; we can ignore it for most analysis purposes.

How is this diagram useful for analysis purposes? It tells us that we can get a rather good approximation of reality by using the double pendulum model and tweaking the shoulder torque profile (that is, shoulder torque vs time) to provide the actual accelerating force that the hands see.

How can we use this to analyze non-circular swings? From photographs or slow-motion measurements of the swing, plot the path of the hands as it varies during the downswing. The key piece of information needed is r(t) , the distance from the path to the center of rotation, as it varies over time. Using r(t), calculate the shoulder torque as a function of time that would give the same accelerating force Fa if the model were a conventional double pendulum of fixed radius R. Then just run the double-pendulum model using the newly calculated shoulder torque, and you will get a behavior that mirrors the noncircular swing, especially in the vicinity of impact. Actually, that is an oversimplification. It would work if all the mass of the arms, hands, and club were accelerating linearly, so we could apply simple F=ma. Since it is rotating, we have to calculate the varying moment of inertia of the arms, hands, and club as r(t) varies. That sounds complicated, but it isn't bad at all in practice. The diagram changes to the one at the left, stressing moment of inertia instead of forces. If you are interested in more detail of the altered model, including an example of its use, you can get it in

my article on the right-side swing.

But there is a constraint on where you can use this modification of the model. The "strobe" diagram on the right is adapted from the SwingPerfect computer program. The circular path of the hands is clearly apparent as the collection of green and red dots, representing the hands at each "snapshot". I have modified SwingPerfect's diagram to color-code the dots: green while the initial wrist-cock angle is still intact, and red once the club swings out and releases the wrist cock. As long as the wrist cock angle is not changing (green dots), our modeling is quite good. But, once centrifugal force starts to release the clubhead (red dots), accuracy depends on the hands being on the circular path. There are a few reasons:

The double-pendulum model reflects that most of the clubhead speed is due to the release of the wrist cock transferring energy from the hands and arms to the clubhead. How much energy is transferred depends on the curvature of the path of the hands during the release. If we change the path of the hands, we will get a different clubhead speed. So the path of the hands once the wrist angle starts increaing must be the circular path being calculated by the model. During release (once the wrist angle starts increasing), tension in club's shaft is exerting a force on the hands that slows them down. Notice that means the force is one that opposes the shoulder torque, which is accelerating the hands. But opposing shoulder torque is what moment of inertia does. So, since we are varying the model by playing

with moment of inertia, our formula for MOI would need to get a lot more complicated during release in order to reflect this force. Better we get to the circular path before release, so we can use the proper double-pendulum model and a relatively simple formula for moment of inertia.

So we have a condition for the altered model to work: By the time the wrist cock reduces significantly from its original angle, the hands must have reached the circular path assumed by the model. That is true for many interesting variants of the swing. In fact, the so-called "standard swing", where there is no wrist torque applied except to keep the club from falling in to the center of the swing, has essentially no change in wrist cock angle for about 60% of the downswing time.

So What Is Double Pendulum NOT Good For? It would be a bit much to assume that as simple a model as the double pendulum is the way to answer all possible questions about the swing. And indeed it isn't. While it is a very good predictor of clubhead speed and loft at impact, there are some things it can't begin to handle. Here are a few examples of things it cannot tell you:

As noted above, any out-of-plane motion of the club. Left-right face angle at any point in the swing, including impact. Clubhead path, or swing plane. How any action below the shoulders affects the shoulder torque. For example, how about the "X-Factor", the angle between hips and shoulders during the downswing. That has to be analyzed separately. The double pendulum model uses shoulder torque as an input to the analysis, not an output from it. Similarly for questions like whether stack-and-tilt is a good idea, or whether a constant frontal spine angle is a good or bad thing. Both of those are inputs to the double pendulum model, not outputs. If you can calculate their effect on shoulder and wrist torque -- which the model doesn't tell you how to do -- then the double-pendulum model can tell you how those torques translate into clubhead speed and effective loft.

It could conceivably calculate in-plane shaft bend, though most existing simulation programs don't produce the information needed to do this calculation.

Bottom line: The double pendulum is an effective model for what goes on:

In the swing plane Between the neck/shoulders and the clubhead Including the right arm.

But it does not deal with issues like out-of-plane motion, clubface direction, or how the body and legs contribute to the swing.

Ben Hogan, Lee Comeaux, and the Right Hand Hit Dave Tutelman -- November 27, 2010

Ben Hogan advocated hitting with the right hand (even both hands) at impact. Recently, Texas instructor Lee Comeaux has been teaching his students a "right-hand slap", which seems very Hogan-esque. One of his students, who has seen marked improvement in his distance, asked me to investigate the physical basis of the improvement. My brief study concludes that it's not what the proponents believe, and the extra distance might not have any explanation in simple physics. Hitting with the hands at impact produces no additional clubhead speed -- none. That might not prevent it from being effective instruction, though.

A Little Background Before I start, let me apologize to my left-handed readers. Hogan and Comeaux use 'right' and 'left' in describing a right-handed swing. I will follow suit. You lefties will have to mentally transpose. (It's not that I don't sympathize; my own son is left-handed. But trying to accommodate it in the wording gets very clumsy.) Hogan and Comeaux

Ben Hogan made no bones about it! He believed in hitting the ball with the hands at the point of impact. Quoting from his classic book "Five Lessons: The Modern Fundamentals of Golf": "Let us study the correct motion of the right arm and hand in the impact area.... On a full shot you want to hit the ball as hard as you can with your right hand. But this is only half the story. HIT THE BALL AS HARD AS YOU CAN WITH BOTH HANDS. The left is a power hand also. If you hit hard with only the right and let the left go to sleep, you will not only lose much valuable power, you also will run into all the errors that result when the right hand overpowers the left."

About a month ago (early November 2010), I got email from Doug "Rock" Burke, telling me about Leecommotion, his name for the swing taught by Lee Comeaux. Lee is a teaching pro, and he teaches a motion that he calls a "shovel move" or "right hand slap". You can see a video of his swing on

YouTube

by

clicking

on

the

thumbnail

to

the

right.

Rock is a low-handicap golfer who reports that the Leecommotion has increased his distance significantly. "I have picked up 30 yards on my drives and 1 club on my irons. It feels a lot more effortless also." It wasn't a trivial change; he practiced it for weeks before he got better. At the time of this writing, he has been at it for 12 weeks, still practicing, and feels it has made a major improvement in his game. My initial impression of Leecommotion is that its essence is the right-hand hit in the impact area, and Rock and I agreed that I should start by studying that. It turns out to be a lot more than that, and I'll do a separate study for "the whole package". But it is worth mentioning in this study, since Hogan's book is not the only school of thought that feels hand action near impact will increase clubhead speed. Rock asked me to explain the benefit in terms of physics. It's one thing to make an improvement; it's quite another to understand why -- and Rock wanted to understand why. This article applies more to Hogan's "hit the ball hard with both hands" Than to Comeaux's "right hand slap", but was mostly motivated by Rock's question. Golf Swing Physics

Before we go on to analyze the slap swing, let's review what we know about the physics of the swing. The illustrations are taken from my tutorial article on golf physics. For more detail than is given in that article, see Rod White's article. For more detail on the mathematical model used to analyze the swing, see my article on the double pendulum. The golf swing is analyzed by physicists as a double pendulum. The inner member of the double pendulum is the triangle of the shoulders and the two arms (green in the diagrams). The outer member is the club itself (black in the diagrams). The variables in the analysis are the lengths of the two arms, and the forces on them. Well, not exactly forces; they are turning forces, more precisely known as "torques". (1) The body rotation exerts a torque on the shoulders, which turns the shoulders and arms. (2) The forearms, wrists, and hands might exert a torque on the club via the grip. If you are not certain you understand the model, its applicability is discussed in more detail in another of my articles. In particular, you may be concerned

that the role of the hips or the right arm might not be properly modeled if this is my tool for analysis. You would be right to be concerned. But whether that concern will doom the analysis is a different story. So understand that article before you dismiss the results out of hand. The most notable of the analyses using the double pendulum model is that by Ted Jorgensen ("The Physics of Golf"). He, and just about every other engineer or physicist who has studied the swing, has concluded that the basic efficient swing is one that depends completely on body rotation (shoulder torque), and lets the wrists hinge freely (that is, zero wrist torque). In other words, you get a very efficient swing from turning the body and allowing the inertial forces on the club to cause the release. (Those inertial forces are usually referred to as centrifugal force. This is a vague and undisciplined use of the language, but we'll stick with it unless the distinction needs to be made for pragmatic reasons.) Hitting with the hands constitutes the addition of wrist torque late in the downswing. This is the basis of our physical analysis of the Hogan and Comeaux swings.

What It Might do I used three distinct approaches to determining the value of adding wrist torque late in the downswing: 1. Stand on the shoulders of giants. Specifically, see what people concluded who had seriously looked at the problem before. 2. Try it myself. Of course, this will only be anecdotal -- hardly a controlled scientific experiment. But it will give me a feel for how the swing works, an insight into the analytic results, and ideas for what to try next. 3. Do the math. I used a computer program that simulates the swing, where the torque profiles can be varied over the downswing. Shoulders Of Giants I admit that I started this investigation with a lot of skepticism because I had read what several "giants" had to say about it.

Let's start with Rod White's analysis. He devotes a subchapter to what happens if you add wrist torque to the standard swing. The graph he presents is unequivocal. Positive wrist torque (torque to uncock the wrists) results in less distance, not more. It may seem paradoxical, but that's what physics says. It is worth noting that negative wrist torque (to hold the wrist cock angle) results in more clubhead speed and more distance. Tom Wishon has been the Chief Technical Officer of Golfsmith, and now has his own golf club component company. His company is recognized as an innovator, and he is the world's foremost advocate of custom fitted golf clubs. In his book "Common Sense Clubfitting" (chapter five on shaft fitting), he writes: "The condition of the shaft being slightly bent backward with the head lagging behind the shaft is very rare in the game. This is because the swing skill plus strength that is required to maintain radial acceleration and the wrist-cock angle until very late in the downswing is such that very, very, VERY few golfers can do this. Far more common are the conditions in which the shaft arrives at impact either straight or slightly bent/curved forward with the head in front of the forward curve of the shaft." In other words, it takes more athletic ability than almost anybody possesses to apply a positive wrist torque late enough in the swing to be helpful. (If I may add my own opinion, not Wishon's words: It isn't just strength; it's also speed. Late in the swing, centrifugal force is whipping the clubhead toward impact. The wrist-cock angle is being dragged out very fast, so it takes a lot of hand speed to even keep up, much less help it along.) One other thing he mentions: we can tell if the wrist torque is negative or positive on the basis of shaft bend. That is important, and we will come back to it later.

Jorgensen's book is a bit more optimistic. In his chapter 4, some of the points he makes are: o It is always a bad idea to apply positive torque early in the downswing. o It is always a bad idea to apply positive torque throughout the downswing. (We knew that already from Rod White.) o It is generally helpful to apply negative torque (that is, hold the lag angle, retard release) throughout the downswing. (Also Rod White.)

What is new and interesting is his analysis of adding positive torque late in the downswing: "Let us look at such a helping action of the wrists late in the downswing. Extensive calculations with the standard swing modified by a constant helping torque of 2 ft-lbs by the wrists late in the downswing show that indeed such a helping torque does produce an increase in clubhead speed at impact. The maximum increase in clubhead speed at impact comes with the torque starting about seven hundredths of a second before the ball is hit. The resulting clubhead speed was found to be only about 0.7% greater than that of the standard swing. This constant torque acting for a longer time or for a shorter time at the end of the downswing produces clubhead speeds less than the maximum." In other words, there is room for a properly executed push or slap to increase power. But it is not much of an increase, and it has to be timed very precisely. Not extremely promising, but not a closed door either. The other giants didn't offer any hope at all.

So we have found a general consensus among those who have investigated the swing mathematically: 1. A strong shoulder turn increases clubhead speed. (The model does not say whether the turn should be a product of the shoulders, the torso, the hips, the legs,... Just that a strong turn powers the swing well.) 2. Wrist torque that helps release usually winds up hurting clubhead speed at impact. You may get higher clubhead speeds earlier in the downswing, but those speeds don't hit the ball; by the time of impact, the clubhead speed is lower than with no wrist torque at all. 3. Wrist torque that retards or delays release usually winds up increasing clubhead speed at impact. #2 and #3 are recognized as paradoxical, but they work time and time again on the computer -- and also on the driving range. 4. When it comes to a "pulse" of helping torque just before impact, there is no consensus. Jorgensen found there is some small advantage to be gained. Wishon believes it is nearly impossible for a real, live human being to achieve that gain. And White didn't even look at the question. (We will look at it a bit later.)

Before we move on, let's take a look at why helping (positive) wrist torque hurts the ultimate clubhead speed, while retarding (negative) torque increases clubhead speed. Here are three "strobe view" simulations of the swing for us to compare. When viewing them, bear in mind that the release is powered, at least in part, by centrifugal force -- and centrifugal clubhead acceleration depends on:

Rotational speed. The faster the golfer rotates, the more the force. Wrist cock angle. The more acute the wrist cock angle, the more the centrifugal force is contributing to clubhead speed, by throwing the club outward. Conversely, the more the wrist cock is already released, the less the clubhead can be thrown outward (accelerated) by centrifugal force.

Of course, body rotation is faster later in the downswing. So we can get the best acceleration from centrifugal force if we hold the wrist cock angle until late in the downswing when fast rotation does the most good. Let's look at the pictures:

No wrist torque (the Positive (helping) Negative (retarding) "standard swing") wrist torque wrist torque

Release is driven only by centrifugal force in this

Maximum acceleration is early, 80msec

Maximum acceleration is late, only

swing.. Maximum acceleration is 60msec before impact. (The reddot clubhead is the point of maximum acceleration.)

before impact. Very little wrist cock is left late in the downswing, stealing centrifugal acceleration. The clubhead positions are closer together at impact, indicating less clubhead speed.

40msec before impact. Wrists stay cocked late into the downswing, encouraging centrifugal acceleration. The clubhead positions are farther apart, indicating more clubhead speed.

(The pictures were generated by the SwingPerfect computer program. We'll be using it more in a later section.) One interesting observation: Jorgensen found that "late wrist torque" should be applied to a "standard swing" beginning "seven hundredths of a second" (70msec) before impact, and continuing through impact. That corresponds almost exactly to the point of maximum acceleration. So his conclusion might be restated, "As long as inertial acceleration is increasing, don't mess with it. Once it starts to fall off, you might help it along with wrist torque." That actually makes sense! Tried It I took it to the practice field and then the course. It integrated more quickly and easily than I expected. I've been using it for a couple of weeks, and every round is the same experience. I hit a bunch of really bad shots early in the round, then find my [new?] swing and hit some very good shots. Some shots are significantly longer than I am used to; I have to believe the new swing has something to do with it. Specific things I have observed while integrating a push/slap.

It is not a silver bullet. It only works if my fundamental swing is sound. Any faults in the basic swing seem to be magnified by the slap. If the slap is introduced early, some very bad results occur. The most common and worst result is a seriously left hit. The ball flight indicates

a closed face. This makes some sense, since the slap includes a rotation of the club. Early slap means early rotation, which means rotated past square at impact. Even many of the good impacts were pulls or even pull-hooks. That means a closed face at impact. I started by thinking of it as a slap or throw with the right hand. Lee Comeaux has a video that describes the move as, "Right hand closest to the body starting the downswing. Right hand farthest from the body through impact." This was a reasonable visualization, and did help. Practicing with it for a while, my body was telling me that an energetic straightening of the right arm was the real driving force behind the motion. A little thought told me that both motions had the same effect on the double pendulum model; they added positive wrist torque. (See my article on the double pendulum and the right arm to see why.)

Every round I have played since adopting the motion has included at least one shot that I didn't know I had in me. So it is doing something good. As noted before, this report is anecdotal, not scientific. But it is encouraging, and adds some urgency to find out why it works. Some possible explanations:

Is it physics? We'll do some modeling and other investigation to find out why. Is it biomechanics or kinesiology? Sometimes a swing key has no good explanation in physics, but encourages other good things to happen in the swing. One example: "Accelerate through the ball." Physics says that acceleration makes no difference once the clubface reaches the ball. But experience shows that a body that expects to accelerate through the ball does a bunch of other things right. Is it the placebo effect -- or, more accurately, the "Hawthorne Effect"? This ain't just physics; it's human experience. Note that we need to treat any swing experiments as human factors experiments, not physics experiments. The Hawthorne effect comes from an experiment by Western Electric in the 1920s, where they changed things like the lighting in the Hawthorne, Illinois factory to see what effect it had on productivity. Turns out anything they did improved productivity. More light? Productivity went up. Less light? Productivity went up. The fact that they were changing things and watching results was more important than exactly what it was they were changing. The "takeaway" for golf: Consistent attention to one thing -- almost anything, as long as it's not actually detrimental -- is going to help your golf performance. It's called disciplined practice.

Do The Math OK, time to do the math. Fortunately, I don't have to do it by hand. I have Max Dupilka's SwingPerfect program, which simulates Jorgensen's model of the swing. You can even control the profile of the torques, varying it over the duration of the swing -- both shoulder torque and wrist torque. We are going to use it to see what effect we get from a helping wrist torque late in the downswing.

Here is a screenshot of what SwingPerfect looks like on my computer. It shows both the main window and the controls to set the wrist torque by 20millisecond intervals during the downswing. There are also controls for the shoulder torque during the downswing and lots of other physical parameters: the golf club dimensions and weights, the golfer's relevant dimensions and weight, and properties of the computation like how frequently to calculate a new point in the downswing. Note for the use of SwingPerfect

If you're not going to actually do studies using SwingPerfect, then you can skip this note altogether. Most studies that are good golf science will be variational studies. That is, you are not looking for absolute numbers but rather, "What happens when I change this parameter by that amount?" So want to set it up before the variations are introduced so that:

The swing is as realistic as possible for the class of player you are trying to simulate. Small variations in parameters give small variations in the answers, related only to your parameter variations and not to any artifact of the way you do simulation.

(1) The most important hint is to reduce the time interval to the minimum the program supports, a half millisecond. This greatly improves the accuracy, as well as the stability of the answer in the face of small variations. It will not give as pretty a display, because the strobes are only a half-millisecond apart, but the numerical answers are much better. (2) In the earlier pictures illustrating centrifugal acceleration, I ran the swing without "Lateral Acceleration" (see Physical Parameters, page 3, in the program) to keep the diagram simple. But a more realistic swing would have lateral acceleration turned on, and that is how I did the numerical studies that follow. (3) For most of the parameters, I used the default values. I changed the club parameters to reflect how drivers are different in 2010 from what they were in 1999 when Max wrote the program. I used club length = 45", head weight = 200g, and shaft weight = 65g. I also made sure the initial arm angle and wrist cock were 180 and 90, which is about what most decent golfers do. (4) One final note: Jorgensen uses the opposite sense of plus and minus for wrist torque. SwingPerfect follows the Jorgensen definition. I chose the "wrong" sense for my narrative in this article. Mine is "wrong" to do the math, but makes more sense when talking to golfers who aren't going to look at the differential equations. So, if you're using SwingPerfect, what I call "positive torque" is minus, and vice versa.

OK, let's compute the effect of a swing with no wrist torque until very late in the downswing; Jorgensen says the last 70 milliseconds helps, and any earlier hurts. So we will try swings with zero wrist torque (the "standard swing"), and a few different helping torques for the last 70msec and the last 110msec. The torques we will try are:

2 foot-pounds. Jorgensen used this number, and thought it was quite a lot. I'm not sure it's so big, but perhaps he's right. 5.7 foot-pounds. Rod White thought that 10% of the shoulder torque was kind of an upper limit on the wrist torque a golfer could apply. I tend to agree. The default shoulder torque for SwingPerfect is 57 footpounds, and 10% of that is 5.7 foot-pounds. 12 foot-pounds. Suppose the golfer were actually able to convert all the shoulder torque into arm-piston force (instead of triangle-turning force). This is clearly an upper bound on the wrist torque, and unrealistically high. But let's compute it just to see what could happen. I used a shoulder width of 24" (exactly 2 feet) and a maximum distance between the hand pressure points (using a 10-finger grip) of 5". 57 ft-lb divided by 2ft gives 28.5 pounds of force in the arm piston. Applied over a 5" grip spread, this is 12 foot-pounds of torque.

Here is a chart of the results. The entries in the chart are:


Clubhead speed, in mph. (Black on white) Clubhead speed compared to standard swing, in percent. (Red on white) Wrist cock angle at impact, in degrees. (Black on yellow) Helping starting 70msec impact 109.3mph (+0.6%) 4.8 110.3mph (+1.5%) torque Helping torque starting before 110msec before impact 108.4mph 0.3%) 3.0 108.3mph 0.4%) ((-

Zero wrist Size of helping torque ("Standard torque Swing") 108.7mph 8.8 5.7 (White) ft-lb 108.7mph

2 (Jorgensen)

ft-lb

8.8 108.7mph 8.8

-1.6 113.3mph (+4.2%) -14.5

-7.5 110.8mph (+1.9%) -28.6

12 ft-lb (Full arm piston)

Before you talk about how much there is to be gained at 12 foot-pounds, let me remind you that 12 ft-lbs is a theoretical upper bound. You're not going to get anywhere close to this number. It is based on converting all the shoulder torque into wrist torque. I went back and used SwingPerfect to simulate that more accurately; I reduced the shoulder torque to zero when the wrist torque kicked in. When you do that, the clubhead speed drops to 97mph, and the wrist is cupped about 60 at impact. That's a recipe for total disaster; don't even think about it. So lets consider the first two rows only, to see what might happen with added wrist torque. First clubhead speed: There's perhaps a couple of miles per hour of clubhead speed to be gained. That translates into no more than five extra yards -- certainly not enough to explain the 30-yard gains Rock reports. Not only is the gain pretty small, but the slap has to be timed very precisely! If you start the slap 40msec earlier than you should, you lose the whole gain and then some. Next wrist cock angle: The standard swing has a slightly bowed wrist at impact, with 8.8 of wrist cock angle remaining to be released. As you crank in helping torque, you decrease the wrist cock angle. If you crank enough wrist torque to make a noticeable difference in distance, the wrist angle at impact becomes somewhat cupped. Because this change is due to a rotation of the forearms, the result is a clubface pointing more to the pull-hook side than usual. Whether it is a pull-hook or just a pull depends on what adding the torque does to your clubhead path; that's swing mechanics that the double pendulum does not address. But there is no doubt that the ball will go left of where it would go with the standard swing, unless you do something else to counteract it. And indeed I experienced that in practice.

I did one more mini-study with SwingPerfect. Because of things discovered later in this article, I wanted to answer the question: Suppose the golfer is exercising retarding torque for most of the downswing. We know this produces more clubhead speed. But... Would releasing this negative torque just before impact increase the clubhead speed further? This would appear to be reasonable, in light of the numbers above, so let's check it out: Case Standard (no wrist torque) 2 ft-lb torque of negative Swing Clubhead Speed Wrist impact 8.8 Cock at

108.8mph

wrist 114.5mph (+5.2%)

17.6

2 ft-lb of negative wrist 114.7mph torque, (+5.4%) released to 0 at 70msec 2 ft-lb of negative wrist 113.4mph torque, (+4.2%) released to 0 at 110msec

14.2

12.3

This tells us that negative wrist torque is a good thing for clubhead speed. By itself, it gives more than a 5% increase in clubhead speed, enough for a gain of about 17 yards. We knew it would give us an improvement, but that's big. Releasing that negative wrist torque 70 milliseconds before impact buys us a tiny bit more clubhead speed -- but not enough for even one more yard. Definitely not worth learning. And if you do your release too early, you lose that gain and a lot more. What It Doesn't Do

So far, we have found out that a right hand hit or slap is modeled for analysis purposes as a wrist torque introduced very late in the downswing. If it is applied for the last 70 milliseconds of the downswing, we may see a small increase in clubhead speed. Not enough for an extra ten yards, but probably a measurable increase. It must be applied at a precise time; much earlier than 70 msec does more harm than good, and much later just doesn't help. So is that the mechanism for the increased distance experienced by users of the right-hand hit? Is there an objective way to find out? It would almost seem foolish to ask this question. There is no doubt in the minds of those who swing with a slap that they are pushing the clubhead through impact. But we cannot dismiss lightly Tom Wishon's admonition that almost nobody can keep up with the releasing clubhead well enough to supply additional acceleration to the clubhead with the hands. So let's see if we can verify that the positive wrist torque is actually being applied to the grip. Shaft Bend It turns out there is a way to find out whether the slap actually causes a positive wrist torque. We just look at shaft bend. But first we have to understand shaft bend. The shaft is a spring that transmits wrist torque from the grip to the clubhead. If you bend the shaft, it transmits a force to the other end that tries to move the object there. Specifically, it tries to move it in a direction to straighten the shaft. That's what any spring does -- apply a force so it can get back to its "resting shape". (Just a reminder: we are bending the shaft in flex, not around its axis -- what shaft specs refer to as "torque".) During the downswing, three conditions can prevail: a straight shaft, a shaft bent backward, and a shaft bent forward. All three show up at one point or another. Let's look at the three conditions:

Straight shaft - If the shaft is straight, that means it is conveying no lateral force to the clubhead. The hands are not applying a wrist torque to the grip, neither to accelerate nor decelerated the clubhead. For the standard swing with freely hinging wrists, the shaft would find itself straight once the wrist cock angle "left the stop". So why is there shaft bend in almost every swing we see, and long after the wrist cock starts to release? The most important reasons are:

Not many players keep their wrists as frictionless hinges. They will exert some torque, whether they are trying or not. Because the center of gravity of the clubhead is not in line with the centerline of the shaft, the centrifugal tug of the clubhead will bend the shaft a little bit. There is certainly bending early in the downswing, where the "stop" of the standard swing is applying torque. Some "rebound" from this initial bend may occur. But, if the wrists were truly hinging, then this effect would be small and short-lived; it would be long gone well before impact.

Bent backward - If wrist torque is accelerating the clubhead toward the ball, then the shaft will be bent backward. That is the only way that the spring of the shaft can apply a helping force to the clubhead. Remember, the force will be applied in a direction to straighten the shaft. So a shaft bent backward -- the clubhead trailing the grip -- is a sure indicator of a positive or helping wrist torque.

Bent forward - By the same reasoning, a forward shaft bend shows a negative, or retarding, wrist torque. Again, the force applied by the shaft is in the direction to straighten the shaft. The force applied to the clubhead by a forward-bending shaft is slowing the release of the clubhead. A little bit of forward bend in the vicinity of impact may be due to the clubhead's center of gravity. But that will seldom be as much as an inch of bend, and most swings have a lot more forward bend than that coming into impact. How Right-Hand Slappers Bend The Shaft This is all very interesting, but how can we use it to learn about the righthand slap? Let's go to the videotape! (Always wanted to say that, even though these days it's an MPEG file or YouTube page, not a tape at all.) I have face-on videos of several golfers that purport to do a right-hand slap, including Lee Comeaux and Ben Hogan. Let's look at freeze-frames near impact. We'll see how the shaft is bent, and deduce from that whether they

are

actually

applying

helping

torque.

Let's start with some frames of Lee from the video we pointed to earlier.

Lee starts down with the wrists well- The backward bend of the shaft at this point of the swing is typical, and cocked. represents the "stop" that keeps the wrist cock from collapsing inward.

Here centrifugal force has taken over, and is trying to uncock the wrists. Lee claims to be doing a right-hand push at this point, but the shaft bend clearly indicates that the hands are actually opposing the uncocking, not causing it.

At this point, the ball is gone. With most swings I've observed, the shaft would be bent backward, at least near the tip, in reaction to impact with the ball. The fact that Lee's shaft is still bent well forward suggests that the clubhead came into impact bent well forward -- with the clubhead's considerable momentum still pulling the hands through -- and impact caused only partial recovery.

In Hogan's peak years, slow-motion video was not nearly as good as today. In spite of the ample footage of Hogan available, very little is of high enough quality to look at shaft bend. But I did find one clip on YouTube good enough to try to draw some conclusions. Here are a couple of frames from that clip.

Fairly late in the downswing, the bright steel shaft is clearly bent forward. If Hogan was able to "hit the ball hard with both hands", it kicked in later than this. Our analysis says that is too late to help clubhead speed measurably.

It is hard to see the shaft in this frame; remember what I said about video quality from Hogan's heyday. But, if you can make out the blur that is the shaft, you can see it is still bent forward. No hand-hit working here, only a few milliseconds before impact.

If you want a couple more pictures of shaft bend, here are two of Lee Comeaux's students, Rock (left) and Karl. Again, you can see that the shaft is bent forward. In Rock's case, the picture was a couple of milliseconds before impact. Any hand hit he can introduce after this does not have enough time to change the clubhead speed. Karl has already hit the ball, though less than 2msec before this frame was snapped. The shaft was deeply bent forward just before impact. You can see a "deflection wave" of back-bending starting up the shaft, caused by impact with the ball, but the shaft was bent forward in the extreme just before impact. In order to develop a big forward bend like this, the clubhead had to be pulling the hands through release; the hands could not have been pushing the clubhead. Both of these are very good swings, and have a lot of power. Both these golfers credit Lee Comeaux with adding power to their game. But we see

here that, whatever is causing the improvement, it is not because the right hand is pushing the clubhead through impact. Physics says that just isn't happening.

Photo from Golf.com

Finally, for those who feel that Long Drive

competitions change the rules, physics included, here's Jamie Sadlowski just before impact. (Jamie won the ReMax World Championship the past two years running, and won handily.) Yes, his shaft is bent way forward, telling us he gets his distance from his huge body turn in the backswing, inhuman wrist cock, and holding that wrist cock very late into his downswing -- and not from forearm or hand strength driving the clubhead through the ball. Likewise for Tiger Woods, as if there were any doubt. Why? Ben Hogan is convinced that hitting the ball with his hands is a big source of his power. Rock and, to a lesser extent, Lee Comeaux lean toward that opinion as well. They have worked and practiced to make it happen. They can feel it happening. But the physics and the videos say that the clubhead is pulling the hands into the ball; their hand-hit is not accelerating the clubhead. What is going on here? My take on Tom Wishon's statement is the key. If you try to add wrist torque late enough in the swing to help (the last 70msec before impact), the hands just can't keep up. Centrifugal acceleration is turning the hands so fast that the wrist muscles and right-arm piston can't turn the hands as fast as the club is already turning them. And if the muscles can't make the hands exceed what the club is doing, then they can't add any acceleration to the clubhead.

In order to test this theory, we need to measure the speed the hands can rotate a club without the load of a club's moment of inertia. Today's digital cameras make video the most convenient way to measure such speeds. So I made a quick video. Click on the photo at the right to view it. I took a short length of PVC plastic pipe, about the diameter of a golf grip. I did a 90 back-and-forth whipping of this baton as fast as I could, simulating the attempt of the hands to uncock and recock the wrists. Because the baton is short and light, it offers little resistance to the turning, so I could measure what the hands would do moving unconstrained at full speed. (An engineer would say that the baton's moment of inertia is negligible compared to that of a golf club.) I examined the video in a movie editor that could identify frame times to within 10msec. (Actually that is more precise than the video itself; the interframe period of my camera is 33msec, so I would not trust an answer to be more precise than that.) What I discovered was that it took at least 150msec to make the move of a nearly 90 release. With a ten-finger grip, it was more like 200msec. True, others may be able to make a quicker move than I, and perhaps the back-and-forth is not as fast as a single forward slap (though I suspect it is). So we may do better, but we are not likely to see an improvement to twice that speed. Why do I mention twice the speed? Let's remember that, when we looked at the "standard swing" the same change of wrist angle occurred in just

70msec. That's just inertial acceleration, no wrist torque (no hand hit) at all. So the hands would have to move more than twice the speed they do, just to keep up with what the clubhead is doing to pull the hands around. In order to actually push the clubhead (rather than be pulled by it) the hands would have to go even faster than that. Conclusion: the clubhead is pulling the hands around at twice the speed the hands can move under their own muscle power. So the hands just can't keep up well enough to apply a hit. What It Probably Does Expecting a right-hand slap or late push to give more distance does not have any basis in physics. We know that now. But there is a substantial body of anecdotal evidence that says it works. Most notably, Hogan won a lot of tournaments with a swing intended to do that (even if it in fact did not do it). Why does it work? We have ruled out physics. I think we can rule out the Hawthorne effect. True, a lot of practice of any swing that is not seriously flawed is going to improve your game. But I, for one, saw results in the first round I played with it, after only two practice sessions. So something is going on there besides just practice. Let's mention some of the possibilities, without evaluating them -- yet. I am beginning a study in conjunction with Lee Comeaux to try to determine what is really going on. 1. Something about the motion encourages some other good habit. This might result in more shoulder torque or a later release, or perhaps something else. 2. Following up on the first point, it might be that focusing on forcing a 90 release just before impact causes the golfer to hold the 90 lag until just before impact. That synchronizes the release to when it should be, and perhaps causes the golfer to exercise negative wrist torque earlier, in order to hold the lag until very late in the downswing. 3. A related possibility: Anybody who has played ball-and-stick sports other than golf has learned to depend on hand action to supply power. You need it in baseball, stickball (I grew up in the Bronx), tennis, etc.

For someone like that, it's very hard to train your body and your subconscious to swing through impact with passive hands (the "standard swing"). If you use a late right-hand slap it might not help, but at least it won't hurt if it's late enough. So the slap swing acts as permission to let the hands get involved, in the least harmful way. If you had been allowing the hands to get in the way by releasing early, then the slap move might cure the problem. If this is the answer, the slap move is not so much a powerful move as an antidote to something you are doing that loses power. 4. An aggressive move through the ball -- like the slap -- might be encouraging a strong, committed follow-through from a golfer who normally gives up the swing after impact. (That is sometimes a problem of mine, but not so much since I started using the slap motion.) 5. Or it might be something completely different. Lee's swing involves a lot more than just the right-hand slap, and those who have been practicing everything Lee teaches report big improvements. I started out suspecting his other advice was in support of the slap, and didn't get very deeply into them. But I have talked with Lee, and there are lots of other things he teaches that could have value in their own right. So maybe one or more of these, and not the right hand slap in the impact area, is responsible for the improvement. In any event, Lee and I are going to see if we can apply physics to other parts of the golf swing that he teaches. Summary From time to time, someone comes along advocating a hit with the hands in the vicinity of impact. Hogan was very explicit about it, and Lee Comeaux is a proponent these days. This study has determined that such a move does not by itself increase clubhead speed. The steps that led us to this conclusion were:

According to physics studies using the double pendulum model, it was determined that small improvements in clubhead speed could indeed be achieved by adding helping wrist torque late in the downswing. But it has to be added about 70 milliseconds before impact; any earlier hurts clubhead speed more than it helps, and any later and the force isn't there long enough to have a significant effect on clubhead speed. These facts have been reported by previous investigators and confirmed by our analysis.

We looked at frames of videos of golfers who claim to use a right-hand hit. Wherever we saw the shaft just before impact, it was always bent forward. That means that, whether the golfer thought so or not, the slap was not applying any accelerating force to the clubhead. The reason would appear to be that the grip is rotating too fast approaching impact, due solely to centrifugal acceleration. The golfer's hands cannot move fast enough to catch up to the grip and apply pressure to it.

Das könnte Ihnen auch gefallen