Dependence of release variables in the shot put

Similar documents
Biomechanical Analysis of a World Record Javelin Throw: A Case Study

arxiv: v2 [physics.pop-ph] 26 Jul 2010

The influence of touchdown parameters on the performance of a high jumper

Preliminary study of Accuracy and reliability of high-speed human-motion tracking using miniature inertial sensors

INTRODUCTION & RECTILINEAR KINEMATICS: CONTINUOUS MOTION

Chapter 2. Kinematics in One Dimension. continued

Coimisiún na Scrúduithe Stáit State Examinations Commission. Leaving Certificate Examination Mathematics

James T. Shipman Jerry D. Wilson Charles A. Higgins, Jr. Omar Torres. Chapter 2 Motion Cengage Learning

Lab 1: Jumping Right In

Components of a Vector

LABORATORY II DESCRIPTION OF MOTION IN TWO DIMENSIONS

Motion in Two or Three Dimensions

Chapter 2 One-Dimensional Kinematics. Copyright 2010 Pearson Education, Inc.

2A/2B BIOMECHANICS 2 nd ed.

2-D Vector Equations have the same form as 1-D Kinematics. f i i

Everything Old Is New Again: Connecting Calculus To Algebra Andrew Freda

Projectile Motion Exercises

(a) On the diagram above, draw an arrow showing the direction of velocity of the projectile at point A.

Chapter 3 Acceleration

Engineering Mechanics Prof. U. S. Dixit Department of Mechanical Engineering Indian Institute of Technology, Guwahati Kinematics

Six Sigma Black Belt Study Guides

ISSUED BY K V - DOWNLOADED FROM KINEMATICS

Review of Engineering Dynamics

Lab 1: Projectiles and Numerical Integration

Figure 2.1 The Inclined Plane

Motion in Two Dimensions. 1.The Position, Velocity, and Acceleration Vectors 2.Two-Dimensional Motion with Constant Acceleration 3.

Uncertainty, Error, and Precision in Quantitative Measurements an Introduction 4.4 cm Experimental error

MOTION IN TWO OR THREE DIMENSIONS

MEAM 211. Project 1. Department of Mechanical Engineering and Applied Mechanics University of Pennsylvania January 19, 2007

PHY 221 Lab 3 Vectors and Motion in 1 and 2 Dimensions

The SuperBall Lab. Objective. Instructions

Excerpts from Project Pathways Student Workbooks

PHYSICS. Hence the velocity of the balloon as seen from the car is m/s towards NW.

Motion of a Point. Figure 1 Dropped vehicle is rectilinear motion with constant acceleration. Figure 2 Time and distance to reach a speed of 6 m/sec

vector of point will be and if the point P is in a space and its coordinates are (x, y, z) then position vector can be expressed as

When we throw a ball :

Chapter 4. Motion in Two Dimensions

Chapter 3 Acceleration

Bell Ringer: What is constant acceleration? What is projectile motion?

Lab 5: Projectile Motion

Introduction to 1-D Motion Distance versus Displacement

In this activity, we explore the application of differential equations to the real world as applied to projectile motion.

Problem: Projectile (CM-1998)

Mechanics. Time (s) Distance (m) Velocity (m/s) Acceleration (m/s 2 ) = + displacement/time.

1-D Motion: Free Falling Objects

Chapter 3 Kinematics in Two Dimensions; Vectors

Adding Vectors in Two Dimensions

Physics 200 Exam 1 Kinematics and Force September 27, 2007

Projectile Motion. Chin- Sung Lin STEM GARAGE SCIENCE PHYSICS

LABORATORY II DESCRIPTION OF MOTION IN TWO DIMENSIONS

Introduction to Mechanics Projectiles

Newton s first law. Projectile Motion. Newton s First Law. Newton s First Law

PS 11 GeneralPhysics I for the Life Sciences

Vectors and 2D Kinematics. AIT AP Physics C

401 Review. 6. Power analysis for one/two-sample hypothesis tests and for correlation analysis.

Vectors. Graphical Method. Graphical Method. SEEMS SIMPLE? = 30.5 m/s. Graphical Method. Graphical Method (TIP TO TAIL) S

Chapter 4. Motion in Two Dimensions

Level 3, Calculus

Motion in 1 Dimension. By Prof. Massimiliano Galeazzi, University of Miami

Thursday 12 June 2014 Afternoon

Faculty of Engineering and Department of Physics Engineering Physics 131 Midterm Examination Monday February 24, 2014; 7:00 pm 8:30 pm

Kinematics 2D ~ Lab. Part 1: Type 1 Projectile Launch. BCLN PHYSICS - Rev. Sept/2011

Chapter 10 Momentum, System of Particles, and Conservation of Momentum

APPLICATIONS OF BIOMECHANICS CINEMATOGRAPHY TO RESEARCH AND COACHING ASPECTS OF THE POLE VAULT

Trigonometry Basics. Which side is opposite? It depends on the angle. θ 2. Y is opposite to θ 1 ; Y is adjacent to θ 2.

1. Joseph runs along a long straight track. The variation of his speed v with time t is shown below.

Mark on the diagram the position of the ball 0.50 s after projection.

Physics 30 Momentum and Impulse Assignment 1 version:0

Chapter 3 Kinematics in Two Dimensions; Vectors

Computer Algebra for Physics Examples

KINEMATICS OF A PARTICLE. Prepared by Engr. John Paul Timola

AP Physics 1 Summer Assignment

Chapter 3. Table of Contents. Section 1 Introduction to Vectors. Section 2 Vector Operations. Section 3 Projectile Motion. Section 4 Relative Motion

2.41 Velocity-Dependent Forces: Fluid Resistance and Terminal Velocity

Chapter 10 Momentum, System of Particles, and Conservation of Momentum

Kinematics in Two-Dimensions

PHY 1114: Physics I. Quick Question 1. Quick Question 2. Quick Question 3. Quick Question 4. Lecture 5: Motion in 2D

Physics 115 Mock Midterm Sunday, October 14, 2018 * 1 pm Room 241 Arts Building *

Problem Solving. Undergraduate Physics

Module 17: Systems, Conservation of Momentum and Center of Mass

Chapter 4. Motion in Two Dimensions

4.1 - Acceleration. What is acceleration?

JAVELIN TRAJECTORY SIMULATION AND ITS USE IN COACHING

Chapter 3 Motion in two or three dimensions

2. Two Dimensional Kinematics

5 Projectile Motion. Projectile motion can be described by the horizontal and vertical components of motion.

y(t) = y 0 t! 1 2 gt 2. With y(t final ) = 0, we can solve this for v 0 : v 0 A ĵ. With A! ĵ =!2 and A! = (2) 2 + (!

"APPENDIX. Properties and Construction of the Root Loci " E-1 K ¼ 0ANDK ¼1POINTS

Principles and Problems. Chapter 6: Motion in Two Dimensions

2009 Assessment Report. Mathematics Level 2

Lecture D2 - Curvilinear Motion. Cartesian Coordinates

New Representation of Bearings in LS-DYNA

Chapter 4. Motion in Two Dimensions. With modifications by Pinkney

Energy and Power in (Sports) Biomechanics. Objectives. Contents

Vocabulary Preview. Oct 21 9:53 AM. Projectile Motion. An object shot through the air is called a projectile.

Projectile Motion. directions simultaneously. deal with is called projectile motion. ! An object may move in both the x and y

Available online at ScienceDirect. Procedia Engineering 147 (2016 )

arxiv: v2 [physics.class-ph] 15 Dec 2016

Chapter 4. Motion in Two Dimensions

Lecture Notes Kinematics Recap 2.4 Acceleration

Transcription:

Journal of Biomechanics 34 (2001) 449 456 Dependence of release variables in the shot put MontHubbard a, *, Neville J. de Mestre b, John Scott c a Department of Mechanical and Aeronautical Engineering, University of California, Davis, CA 95616, USA b School of Information Technology, Bond University, Queensland, Australia c Department of Management Systems, University of Waikato, New Zealand Accepted 30 November 2000 Abstract When the shot is released above a horizontal plane, range from this point depends on release height, speed and angle. Measured distance is the sum of this range and horizontal distance of the release point from the throwing circle edge. Optimal release conditions can be calculated only if the dependence of release velocity on other variables, due to thrower limitations, is known. Experiments on two shot-putters investigated the hypothesis that there are constraint relationships among these four release parameters. A variable scaling scheme, using measurement of impact point and the known magnitude of g, corrected 2D data from one camera for out-of-plane motion and yielded accurate estimates of release parameters. Multivariate regression analyses determined approximate constraint surfaces limiting performance. Achievable release speed decreases with increasing release angle at about 1.7 (m/s)/rad and decreases with increasing release height at about 0.8 (m/s)/m, with only small differences in sensitivities between the throwers. Horizontal release distance also decreases with increasing release angle at about 1.7 m/rad and increases with increasing release height at about 1.3 m/m, again with only small differences between the two throwers. Optimal release conditions producing maximum range for a particular athlete can be determined using similar constraints for that athlete. # 2001 Elsevier Science Ltd. All rights reserved. Keywords: Shot put; Optimization; Release variables; Release constraints 1. Introduction The shot put has been studied extensively, both analytically and experimentally. Theoretical studies and subsequent reviews (Garfoot, 1968; Lichtenberg and Wills, 1978; Zatsiorsky et al., 1981; Trowbridge and Paish, 1981; Burghes etal., 1982; Townend, 1984; Hubbard, 1989; de Mestre, 1990) discuss optimal release angle, the angle between the initial velocity vector and horizontal. Lichtenberg and Wills, neglecting any possible relationship between release angle and release speed, proved that the deviation of the optimal release angle from the obvious value of 458 is due to the fact that the release height is approximately 2 m higher than the level of impact. They calculated the optimal release angle, for a speed producing roughly world record distances of 73 ft, to be y m =42.38, and showed that aerodynamic drag has a negligible (50.18) effecton *Corresponding author. Tel.: +1-530-752-6450; fax: +1-530-752-4158. E-mail address: mhubbard@ucdavis.edu (M. Hubbard). optimal release angle. Curiously, however, there is a large discrepancy between measured release angles (McCoy et al., 1984) and the optimal release angles predicted on their theoretical basis. Film studies by Cureton (1939), Dessurealt (1976), Tsirakos et al. (1995), and McCoy et al. indicate that elite shot putters release the shot at an angle to the horizontal considerably less than 428, and sometimes as low as 298. All previous developments derive the formula for the optimum angle of release assuming independence of the release angle, release heightand release speed, i.e. that there are no constraints among these variables. This assumption of independent velocity does not seem to be physically realistic, however, since the effort needed to project the shot horizontally appears to be much less than that needed to project it vertically (e.g. see the explanation by Zatsiorsky, 1990). Two previous investigations found similar constraint relations between release variables in javelin throwing, Red and Zogaib (1977) using a 1.14 kg ball of similar mass and Viitasalo and Korjus (1988) using a javelin. Both determined that the release speed decreased 0021-9290/01/$ - see front matter # 2001 Elsevier Science Ltd. All rights reserved. PII: S 0021-9290(00)00228-1

450 M. Hubbard et al. / Journal of Biomechanics 34 (2001) 449 456 linearly with release angle, but did not consider other variables or higher-order models. The experiments of Red and Zogaib (1977) were apparently motivated by incorrect theoretical predictions of optimal javelin release angles made by Soong (1975), assuming independence of javelin release speed and angle. Recently Maheras (1995) has undertaken a similar study in the shot put, which developed regression equations for release angle as a function of release velocity and height. In addition there has been brief speculation on the possibility of a similar effect by Tsirakos et al. (1995). In a review of the throws in track and field, Hubbard (1989) noted that, in all throwing events, optimal release conditions (for variables) other than velocity depend crucially on how the maximum achievable release velocity is functionally related to the other release conditions, and this is commented on further in Hubbard (2000). The fundamental hypothesis of this paper is that there is indeed a dependence of achievable release velocity and release horizontal distance on the other release conditions. This is important because this functional dependence may account for the remaining discrepancy between observed and predicted optimal release angles. In addition, in order to train throwers to throw optimally, it is necessary to be able to compute what their desired optimal behavior is. The present paper addresses only the determination of the constraint relations among the four variables; release speed, horizontal distance, release angle and height. 2. Methods Before we consider constraints among the release variables, we firstexamine flight. Consider a shotof mass m released ata heighth with speed v and angle y between the velocity vector and horizontal. Choose the point on the throwing circle from which the measurement of range is made as the origin of a two-dimensional (2-D) coordinate system with x 0 horizontal in the actual throwing direction and y 0 vertical (Fig. 1), so the release pointhas coordinates (d; h). Fig. 1. Shot put trajectory geometry with release variables. Total range R is sum of the flight range (a function of v, y and h) and the horizontal release distance d from the edge of the throwing circle. Numerous previous theoretical studies (e.g. Dyson, 1977; Lichtenberg and Wills, 1978; Townend, 1984; de Mestre, 1990) contain perhaps the most fundamental summary of the flight range, assuming a drag-free model. Only Dyson (1977) and Tabor (1990) included a correction d because the shot is not necessarily released directly above the edge of the throwing circle. Adding flightrange to d yields R ¼ d þ v 2 cos y½sin y þðsin 2 y þ 2gh=v 2 Þ 1=2 Š=g: ð1þ Even for world-class shot putters, including air-drag results in less than 1% correction to the range from Eq. (1). Therefore, we here neglectaerodynamic drag and its effect on range. If v; h and d are independentof release angle, the optimal release angle y m maximizing range (Lichtenberg and Wills, 1978; de Mestre, 1990; Hubbard, 2000) is derived by requiring dr=dy ¼ y, yielding y m ¼ arctan½v=ðv 2 þ 2ghÞ 1=2 Š: ð2þ However, where v; y; h and d are notindependent, Eq. (2) does not apply and their interdependence must be determined before optimal release conditions can be calculated (Hubbard, 1989). The purpose of this paper is to investigate these relationships. Two skilled shot putters from the UCD track-andfield team threw the same shot in 36 trials. One (AJB 106 kg, 1.83 m tall } 22 throws) was in the top 12 nationally and the other (JST 115 kg, 1.91 m tall } 14 throws) was at college level. Each was asked to produce three sets of relatively high initial angle, normal and relatively low initial-angle throws, respectively, all at maximum effort, and each used the standard rotary throwing technique. A 200 Hz video (ExpertVision, Santa Rosa, California) motion analysis system was used to determine initial release angle, speed, heightand horizontal distance for each throw as described below. A straight desired throwing line (the x-axis of a second coordinate system; see Fig. 2) was laid out on the ground bisecting the throwing circle with origin also at the center of the inner board edge. The single camera was levelled and placed ata distance c 0 (approximately 2.5 m) from, and with its axis perpendicular to, this line so that its field of view encompassed release and the beginning of flight (Fig. 2). Initial 2-D calibration of the camera relied on a calibration object in the xy plane. Throwers were instructed to throw so that both the release position and initial velocity vector lay in this plane. Even with assistance and feedback, however, the throwers found it very difficult to (and never did) release the shot exactly in the desired plane or in the correct direction. The implications of this on scaling are discussed further below. Camera data were first processed, using the (naive) 2-D assumption that the throw occurred in the xy plane.

M. Hubbard et al. / Journal of Biomechanics 34 (2001) 449 456 451 Fig. 2. Plan view of release and subsequent trajectory. The camera axis is oriented perpendicular to the desired throwing line x, but the actual trajectory neither begins above, nor does its vertical plane remain parallel to, the desired throwing line. Pixel coordinates were scaled by the constant initial calibration scale factor and shifted to account for the position of the origin. Typical trajectory segments contained 15 20 co-ordinates of the centroid of the circular image over roughly 1-m of flightat0.005 s intervals. Impact coordinates (x f ; z f ) were also measured, allowing calculation of actual range. Parabolic (drag-free) approximations of the trajectory were fitted to the data to yield estimates for x and y positions and velocities at the first frame after release. Predicted range R p was calculated using these release positions and velocities and compared to the measured range R m. Results for (JST) are shown in Fig. 3 (as open squares). Clearly there is a substantial discrepancy. A least-squares regression line containing the origin gives R p ¼ 0:963R m, with nearly 4% average error and substantial maximal deviations of about 10%, showing that the 2-D approximation and the calculated release conditions based on it were not valid. The video data appeared very clean, however, and measurementof landing position and range was reasonably accurate (1 cm). Thus, an alternate scaling method using this data was developed to produce more accurate estimates for release conditions. The basic idea is to include parameters in the model which account for the major error sources due to scaling, and to use video and range data to estimate release conditions and error sources simultaneously, thus removing their effects from the calculated release conditions. This technique is an example of using accurate non-image-related information to correct out-of-plane errors (Sih et al., 2001). Consider Fig. 2. Initial calibration determines a nominal scale factor s 0 (meters per pixel), which applies only to motion in the calibration plane. However, the general throw was neither released in the desired xy plane, nor was its x 0 y 0 flight plane parallel to it. Therefore, the correct scale factor s varies with distance c from the camera throughout flight according to s ¼ cs 0 =c 0 ¼ðc 0 þ z 0 þ v x t tan fþs 0 =c 0 ; ð3þ Fig. 3. Measured ranges versus predicted ranges for JST (open squares) using scaling based on the naive assumption that the throw occurred in the desired throwing plane. Scaling pixel data using a constant distance scale factor yields mean and maximal errors in predicted range of nearly 4% and more than 10% respectively. This motivates the variable scaling procedure discussed in the paper, which achieves rms errors of less than 0.01 m (solid squares). where v x is the x velocity and t is the time since release. Combining parameters yields s ¼ s 1 þ kt; ð4þ where s 1 and k are interpreted as initial scale factor at release and (constant) time rate of change of scale factor throughout flight, respectively. Since we propose to scale the pixel data in a somewhat fluid way, it is important to insure that the results remain reasonable. We define a performance index J that includes two well-known requirements. First, predicted range R p using release conditions determined from the scaled data should accurately match the measured range R m. Second, neglecting aerodynamic drag, the y componentof scaled position should be quadratic in time with its t 2 coefficient g a =2 nearly half the known local value of g=9.7935 m/s 2. The performance index J is written as J ¼ðR m R p Þ 2 þðg a gþ 2 : ð5þ The unknown scaling parameters s 1 and k are equivalent to the two parameters c 0 and z 0 since s 1 ¼ð1 þ z 0 =c 0 Þs 0 and k ¼ v x s 0 tan f=c 0 ; ð6þ ð7þ in which v x is nearly constant and f ¼ tan 1 ððz f z 0 Þ=ðx f x 0 ÞÞ and where (x 0 ; z 0 ) and (x f ; z f ) are the initial and final (impact) positions, respectively. Scaling parameters s 1 and k were determined iteratively for each throw as follows.

452 M. Hubbard et al. / Journal of Biomechanics 34 (2001) 449 456 Algorithm for determination of optimal scaling parameters: 1. Make initial guesses for s 1 and k (s 1 ¼ s 0 and k ¼ 0). 2. Scale camera data with s 1 and k using Eq. (4) to yield 3-D positions versus time. 3. Determine position and velocity at the first frame by fitting a quadratic (drag-free) polynomial in time to this position data. Estimate apparent gravity g a as half the quadratic coefficient of the y regression. 4. Use these initial conditions and Eq. (1) to predict impactpointand range. 5. Calculate the performance index J from Eq. (5) and its gradient with respect to s 1 and k. 6. Use the nonlinear least-squares estimation technique described in Hubbard and Alaways (1989) to calculate perturbations ds 1 and dk to the scaling parameters s 1 and k that decrease the performance index. 7. Stop if corrections ds 1 and dk or the change in J are too small; otherwise go to step 2. Note that in step 4, the predicted range is obtained from a drag-free parabolic approximation. Lichtenberg and Wills (1978) showed that, even with velocities near 15 m/s and flighttimes near 2 s, aerodynamic forces decrease range by only 0.8%. Drag-free approximations applied to very short trajectory segments (Dt=0.1 s) incur much smaller errors (0.03%). Because we are interested in the sensitivity of release conditions, and not the conditions themselves, the drag-free approximation was adequate even in the full trajectory. The optimal scaling technique was implemented using MATLAB. The algorithm converges quadratically and achieves an rms error less than 0.01 m between predicted and measured ranges (solid symbols in Fig. 3). It completely resolves the problem from the native 2-D scaling shown by the open square symbols and also achieves agreementbetween apparentand known values of gravity to better than 0.1%. Release angle y was calculated from y ¼ arctanðv y =v h Þ ð8þ and initial speed v from v ¼ðv 2 h þ v2 y Þ1=2 ; where the horizontal velocity is given by v h ¼ðv 2 x þ v2 z Þ1=2 ð9þ ð10þ and v h and v y were determined as in step 3 of the algorithm. The two variables d and h are the coordinates in the x 0 y 0 plane atrelease. Possible constraint relationships among the four variables v; d; y; and h are the subject of the remainder of the paper. Because of the quadratic dependence of R on v, it is more sensitive to changes in v than to changes in the other two variables y and h. For this reason we choose v as a dependentvariable and y and h as independent variables in one of the constraint relations. Consequently, the remaining variable d is also chosen as a dependentvariable in y and h. Labelling certain variables as independentand others as dependentis somewhatarbitrary, since any constraintrelationships among the variables treats them all in an equal footing. Thus, the terms independent and dependent apply more to the regression model than to the constraint relationships themselves, and should not imply that some variables are more important or fundamental than others. Multivariate polynomial regression analyses were used to elucidate the relationships imbedded in the release data. The regression equations were solved using a MATLAB program. So that the sensitivities obtained would be as meaningful as possible to the reader, the independentvariables in the regressions were chosen as the deviations Dy and Dh from mean values y m and h m for each athlete. Quadratic multiple regressions were performed for each thrower of the form v ¼ C þ C y Dy þ C h Dh þ C yy Dy 2 þ C yh DyDh þ C hh Dh 2 ð11þ with a similar equation for the other dependent variable d. Finally the regression coefficients were tested for statistical significance. The most common null hypothesis H o is that all coefficients except the mean C in a given regression are zero, i.e. that the dependent variable is not related to the independent ones at all (Bethea and Rhinehart, 1991). This hypothesis is rejected using a one-sided F test based on the ratio of the mean square due to regression with p degrees of freedom (MSR) to the mean-squared residual error (MSE). The null hypothesis can be rejected at the 1 a level if the statistic F¼ MSR=MSE >F p;ðn p 1Þ;1 a (Bethea and Rhinehart, 1991, p. 171). In the models above, the number of regression degrees of freedom p ¼ 5. 3. Results Achievable release speed decreases with increasing release angle atabout1.7 (m/s)/rad and decreases with increasing release heightatabout0.8 (m/s)/m, with only small differences in sensitivities between throwers (see top half of Table 1). For each thrower the sensitivities of release speed to y and h are within 3% of 1.68 (m/s)/ rad (speed decreases as angle increases) and within 15% of 0.8 (m/s)/m (speed decreases as heightincreases), respectively. In addition the principal curvatures of the speed angle height constraints for the two athletes are similar in sign and within a factor of two in magnitude. In addition to the regression coefficients, Table 1

M. Hubbard et al. / Journal of Biomechanics 34 (2001) 449 456 453 Table 1 Least-squares regression coefficients for release variable constraints. Achievable release velocity and horizontal distance depend on release angle and release height Name Regression coefficients Means rms error Confidence level C C y C h C yy C yh C hh y m (rad) h m (m) (m/s) 1 a (m) AJB v 11.52 1.73 0.93 13.12 16.2 9.59 0.73 1.97 0.14 0.965 JST v 11.22 1.64 0.68 6.3 5.89 14.64 0.63 2.13 0.12 >0.995 AJB d 0.09 1.81 1.24 0.11 1.5 3.25 0.73 1.97 0.03 >0.995 JST d 0.03 1.64 1.31 6.69 2.14 1.23 0.63 2.13 0.05 >0.995 contains the means of the independent and dependent variables, and the rms errors of the fits. Horizontal release distance also decreases with increasing release angle atabout1.7 m/rad butincreases with increasing release height at about 1.3 m/m, again with only small differences between the two throwers (bottom half of Table 1). Although mean values for horizontal release distance for both throwers were negative (each usually released behind the line) their magnitudes were less than 0.1 m. There are significant sensitivities of d to both y and h, however, which are again nearly equal for each thrower. As with the release speed sensitivities, there are only relatively small differences between the two throwers. The constraints are perhaps more easily visualized graphically. The coefficients from Table 1 were used in Eq. (11) to produce each thrower s constraint surfaces for v ¼ vðy; hþ and d ¼ dðy; hþ. These are illustrated in Figs. 4 and 5, respectively, using the same scales for v, d, y and h. Also, included are the experimental points from which the least-squares fits and coefficients were calculated. A measure of the errors between the data and the fits are depicted by vertical lines connecting the data to the surfaces. The regression coefficients are shown to be statistically significantby the lastcolumn of Table 1. This contains the confidence level 1 a atwhich the null hypothesis H o, that the dependent variable for that row is not related to the independent ones at all, is rejected. In every case 1 a > 0:95. The least statistically significant relationship is that for v for AJB butall others have statistical significance greater than 0.995. 4. Discussion These constraints are inevitably caused by limitations of the geometry and musculoskeletal structure and ability of the thrower. Because different throwers have different heights, strengths, arm lengths and muscular dynamic function (force velocity and force length characteristics), one might expect that such constraint relationships would be complex and different for different throwers. Although we are not able to draw any universal conclusions based only on two throwers, it is interesting that the constraints for the two so physically dissimilar athletes are so nearly equal. The fits to v and d in Figs. 4 and 5 are relatively good in the sense that the rms errors are only about 0.13 m/s and 0.04 m, respectively, roughly 10% of the variations of v and d in the experiments. Maheras (1995) developed regression equations for shot put release angle as a function of velocity and height, and thus his results are difficult to compare with ours. The negative slopes of his velocity-release angle scattergrams, however, range from 2.8 to 5.7 (m/s)/ rad, roughly a factor of two higher than the results in this study, 1.68 (m/s)/rad. The release velocity sensitivity to release angle in the shotputof 1.68 (m/s)/rad is about 4 times smaller than the similar javelin sensitivity of 7.28 (m/s)/rad (Red and Zogaib, 1977, Viitasalo and Korjus, 1988), but the negative signs indicate that in both cases the throwers can throw faster at smaller release angles. The magnitudes of these sensitivities of d to y and h are reasonable based on geometric arguments. Because thrower AJB is 0.08 m shorter than JST and had correspondingly shorter arms, his mean release heights were considerably less (0.16 m) than those of JST. Although the derivative of release d with respect to release h is not the same concept as the variation of horizontal distance with height along a trajectory ðdx=dy¼ cot yþ, their magnitudes should be of the same order. For the overall mean release angle of y ¼ 0:68 rad, cot y ¼ 1:23, which is remarkably close to 1.27, the mean sensitivity of d to h from Table 1. The constraint relations discussed above can be interpreted as approximate limitations on the ability of a thrower to achieve combinations of release speed and horizontal release distance which will increase attainable range. This is because the experiments were conducted with a request for maximal effort by the throwers. Due to the nature of the regression process, the constraint surfaces roughly bisect the data points. The true limiting surfaces (actual boundaries of performance) will be nearly parallel to these but shifted slightly to lie outside

454 M. Hubbard et al. / Journal of Biomechanics 34 (2001) 449 456 Fig. 4. Release velocity constraint surfaces versus release angle y and release height h for two throwers. Surfaces are least-squares quadratic fits of speed data v (plotted as * ) and are portrayed as lines of constant y. Vertical lines denote distance from data * to surface. Rms errors in the surface fits are 0.14 and 0.12 m/s. Achievable release speed decreases with increasing release angle at about 1.7 (m/s)/rad and decreases with increasing release heightatabout0.8 (m/s)/m. all (or possibly on some of) the data. As an approximation we may take the constraint surfaces determined above as adequate representations of the true limiting constraint surfaces. The limitations of the approach we have followed to identify the constraints come from two main factors that prevent their exact determination. Most importantly we have assumed that the throwers were operating at maximal effort. But the throwers are unable to behave like machines and thus are not likely to achieve repeatedly their maximal velocity in every throw, although they might have been trying as hard as possible to do so. Were they able to, all the data points would lie exactly on the actual constraint surfaces. Secondly, even if they were able to do this, ubiquitous measurementerrors would preventprecise determination of the release conditions and hence the constraints. Whatever the constraints are, they will reduce the problem of finding the release conditions for maximum range to that of optimizing a function of only two independentvariables. This is considerably more complicated than the analyses undertaken by most researchers to date (Maheras, 1995). Previous efforts, except Tabor (1990) who assumed that d was constant, have postulated no relationships among v, d, h, and y. They merely soughtthe maximum range for fixed values of v and h with only y varying and assuming d ¼ 0 or another constant. The introduction of the horizontal release distance d in the total range expression for the throw of a shot, and the recognition that both d and the release speed v depend on the other release conditions, y and h, provide a framework for a more accurate optimization of the throw to achieve maximum distance. This optimization depends crucially on the knowledge of the constraints relating the release variables for each athlete. This paper has described experiments, data processing techniques, and multivariate regression analyses used to develop these constraints. In summary, for a given thrower, the shot can be thrown faster when thrown at smaller release angles and when released lower. Achievable release speed decreases with increasing release angle at about 1.7 (m/s)/rad and also decreases with increasing release height at about 0.8 (m/s)/m. Release horizontal distance decreases with increasing release angle atabout1.7 m/rad, butincreases with increasing release heightatabout1.3 m/m. These sensitivities can be used in optimization studies to

M. Hubbard et al. / Journal of Biomechanics 34 (2001) 449 456 455 Fig. 5. Release horizontal position constraint surfaces versus release angle y and release height h for two throwers. Surfaces are least-squares quadratic fits of horizontal position data d (plotted as * ) and are portrayed as lines of constant y. Vertical lines denote distance from data * to surface. Rms errors in the surface fits are 0.03 and 0.05 m. Horizontal release distance also decreases with increasing release angle at about 1.7 m/rad butincreases with increasing release heightatabout1.3 m/m. predict optimal release conditions which may agree more favorably with experiment than in the past. Acknowledgements The authors would like to thank the athletes Allan Babayan and Joey Taylor as well as Daniel Reed and Charles Black for assistance with the experiments. References Bethea, R.M., Rhinehart, R.R., 1991. Applied Engineering Statistics. Marcel Dekker, New York. Burghes, D.N., Huntley, I., McDonald, J., 1982. Applying Mathematics. Ellis Horwood, Chichester. Cureton, T.K., 1939. Elementary principles and techniques of cinematographic analysis. Research Quarterly 10, 3 19. de Mestre, N., 1990. The Mathematics of Projectiles in Sport. Cambridge University Press, Cambridge. Dessurealt, J., 1976. Selected kinetic and kinematic factors involved in shot putting. Ph.D. Thesis, Indiana University. Dyson, G., 1977. The Mechanics of Athletics. Hodder and Stoughton, London. Garfoot, B.P., 1968. Analysis of the trajectory of the shot. Track Technique 32, 1003 1006. Hubbard, M., 1989. The throwing events in track and field. In: Vaughn, C.L. (Ed.), Biomechanics of Sport. CRC Press, Boca Raton, FL, pp. 213 238. Hubbard, M., Alaways, L.W., 1989. Rapid and accurate estimation of release conditions in the javelin throw. Journal of Biomechanics 22 (6 7), 583 595. Hubbard, M., 2000. The flight of sports projectiles. In: Zatsiorsky, V.M. (Ed.), Biomechanics in Sport: Performance Enhancement and Injury Prevention. Blackwell Science, Ltd., Oxford (Chapter 19). Lichtenberg, D.B., Wills, J.G., 1978. Maximizing the range of the shot put. American Journal of Physics 46, 546 549. Maheras, A.V., 1995. The relationship between the angle of release and the velocity of release in the shot-put, and the application of a theoretical model to estimate the optimum angle of release (throwing). Ph.D. Thesis, University of Kansas. McCoy, R.W., Gregor, R.J., Whiting, W.C., Rich, R.G., Ward, P.E., 1984. Kinematic analysis of elite shot putters. Track Technique 90, 2868 2871. Red, W.E., Zogaib, A.J., 1977. Javelin dynamics including body interaction. Journal of Applied Mechanics 44, 496 498. Sih, B.L., Hubbard, M., Williams, K.R., 2001. Correcting out-of-plane errors in two-dimensional imaging using nonimage-related information. Journal of Biomechanics 34, 257 260. Soong, T.C., 1975. The dynamics of javelin throw. Journal of Applied Mechanics 42, 257 262.

456 M. Hubbard et al. / Journal of Biomechanics 34 (2001) 449 456 Tabor, J., 1990. Putting the shot. In: Huntley, I., James, P. (Eds.), Mathematical Modelling } A Source Book of Case Studies. Oxford University Press, Oxford, pp. 117 146. Townend, M.S., 1984. Mathematics in Sport. Ellis Horwood, New York. Trowbridge, E.A., Paish, W., 1981. Mechanics of athletics. The Institute of Mathematics and its Applications 17, 144 146. Tsirakos, D.K., Bartlett, R.M., Kollias, I.A., 1995. A comparative study of the release and temporal characteristics of shot put. Journal of Human MovementStudies 28, 227 242. Viitasalo, J.T., Korjus, T., 1988. On-line measurement of kinematic characteristics for javelin. In: degroot, G., et al. (Ed.), Biomechanics XI-B. Free University Press, Amsterdam, pp. 582 587. Zatsiorsky, V.M., Lanka, G.E., Shalmanov, A.A., 1981. Biomechanical analysis of shot putting technique. Exercise and Sports Science Reviews 9, 353 389. Zatsiorsky, V.M., 1990. The biomechanics of shot putting technique. In: Bruggeman, G.-P., Ruhl, J.K. (Eds.), Techniques in Athletics, Deutsche Sporthochschule, Koln.